<<

Ministry of Natural Resources Community-Level 36 Effects of Climate CLIMATE Change on Ontario’s CHANGE Terrestrial RESEARCH REPORT CCRR-36

Responding to Climate Change Through Partnership Sustainability in a Changing Climate: An Overview of MNR’s Climate Change Strategy (2011-2014)

Climate change will affect all MNR programs and the • Facilitate the development of renewable energy by natural resources for which it has responsibility. This collaborating with other Ministries to promote the val- strategy confirms MNR’s commitment to the Ontario ue of Ontario’s resources as potential green energy government’s climate change initiatives such as the sources, making Crown land available for renewable Go Green Action Plan on Climate Change and out- energy development, and working with proponents lines research and management program priorities to ensure that renewable energy developments are for the 2011-2014 period. consistent with approval requirements and that other Ministry priorities are considered. Theme 1: Understand Climate Change • Provide leadership and support to resource users MNR will gather, manage, and share information and industries to reduce carbon emissions and in- and knowledge about how ecosystem composition, crease carbon storage by undertaking , structure and function – and the people who live and protecting natural heritage areas, exploring oppor- work in them – will be affected by a changing climate. tunities for forest carbon management to increase Strategies: carbon uptake, and promoting the increased use of • Communicate internally and externally to build wood products over energy-intensive, non-renewable awareness of the known and potential impacts of alternatives. climate change and mitigation and adaptation op- • Help resource users and partners participate in a tions available to Ontarians. carbon offset market, by working with our partners • Monitor and assess ecosystem and resource condi- to ensure that a robust trading system is in place tions to manage for climate change in collaboration based on rules established in Ontario (and potentially with other agencies and organizations. in other jurisdictions), continuing to examine the • Undertake and support research designed to mitigation potential of forest carbon management in improve understanding of climate change, including Ontario, and participating in the development of pro- improved temperature and precipitation projections, tocols and policies for forest and land-based carbon ecosystem vulnerability assessments, and im- offset credits. proved models of the carbon budget and ecosys- tem processes in the managed forest, the settled Theme 3: Help Ontarians Adapt landscapes of southern Ontario, and the forests MNR will provide advice and tools and techniques to and wetlands of the Far North. help Ontarians adapt to climate change. Strategies • Transfer science and understanding to decision- include: makers to enhance comprehensive planning and • Maintain and enhance emergency management management in a rapidly changing climate. capability to protect life and property during extreme events such as flooding, drought, blowdown and Theme 2: Mitigate Climate Change wildfire. MNR will reduce greenhouse gas emissions in sup- • Use scenarios and vulnerability analyses to develop port of Ontario’s greenhouse gas emission reduction and employ adaptive solutions to known and emerg- goals. Strategies: ing issues. • Continue to reduce emissions from MNR opera- • Encourage and support industries, resource users tions though vehicle fleet renewal, converting to and communities to adapt, by helping to develop un- other high fuel efficiency/low-emissions equipment, derstanding and capabilities of partners to adapt their demonstrating leadership in energy-efficient facility practices and resource use in a changing climate. development, promoting green building materials • Evaluate and adjust policies and legislation to re- and fostering a green organizational culture. spond to climate change challenges. Community-Level Effects of Climate Change on Ontario’s Terrestrial Biodiversity

Larissa A. Nituch1 and Jeff Bowman1*

1Wildlife Research and Monitoring Section Science and Research Branch Ontario Ministry of Natural Resources Trent University, DNA Building 2140 East Bank Drive Peterborough, ON K9J 7B8

*correspondent: [email protected]

2013

Science and Research Branch • Ontario Ministry of Natural Resources © 2013, Queen’s Printer for Ontario Printed in Ontario, Canada

Single copies of this publication are available from:

Science and Research Branch Ontario Forest Research Institute Ministry of Natural Resources 1235 Queen Street East Sault Ste. Marie, ON Canada P6A 2E5

Telephone: (705) 946-2981 Fax: (705) 946-2030 E-mail: [email protected]

Cette publication hautement spécialisée, Community-level effects of climate change on Ontario’s terrestrial biodiversity n’est disponible qu’en anglais en vertu du Règlement 671/92 qui en exempte l’application de la Loi sur les services en français. Pour obtenir de l’aide en français, veuillez communiquer avec le ministère des Richesses naturelles au [email protected].

This paper contains recycled materials. i

Summary Rapid, anthropogenic climate change has the potential to be a major threat to the biodiversity of terrestrial communities, and is one of the main factors affecting species interactions and ecosystem functioning. Previous reports have described three general mechanisms that can affect species as a result of climate change: demographic, phenological, and genetic, all of which can result in either population expansions or contractions, depending on species-specific responses. In this report, we describe mechanisms that are expected to affect ecological communities, rather than individual species, as a result of climate change.

The effects of climate change on communities and ecosystems are difficult to predict because of complexities and uncertainties associated with biotic interactions. Climate change can significantly affect the genetic composition and structure of communities, and can alter the genetic connectivity among populations, increasing the risk of genetic diversity losses. Climate change typically affects species in communities disproportionately, reducing synchrony and symmetry between interacting species, such as predators and prey. Climate change can also act synergistically with other processes, such as habitat fragmentation, disease, and , to exacerbate the overall effects. Since individual species responses to climate change vary, some will adapt and remain in a community, others will leave a community, and non-native species may join a community. The result is the potential generation of novel biotic communities, referred to as community reassembly. Community reassembly alters community composition and can therefore lead to changes in biodiversity, species interactions, trophic structure, and ecosystem processes. In this report, we discuss the potential community-level effects of climate change on terrestrial ecosystems, with a focus on wildlife, and identify gaps in knowledge. We also make recommendations for associated management consideration, research needs, and adaptation strategies. Résumé Effets au niveau de la communauté du changement climatique sur la biodiversité terrestre de l’Ontario Un changement climatique anthropique rapide est susceptible de menacer sérieusement la biodiversité des communautés terrestres, et c’est un des principaux facteurs influençant l’interaction des espèces et le fonctionnement des écosystèmes. Des rapports antérieurs ont décrit trois mécanismes généraux qui peuvent avoir une incidence sur les espèces en raison des changements climatiques : les mécanismes démographique, phénologique et génétique, qui peuvent tous entraîner un accroissement ou une diminution de la population, selon les réactions propres aux différentes espèces. Dans le présent rapport, nous décrivons des mécanismes qui devraient influer sur des communautés écologiques, plutôt que sur des espèces données, du fait du changement climatique.

Les effets du changement climatique sur les communautés et les écosystèmes sont difficiles à prédire en raison de la complexité et de l’incertitude des interactions biotiques. Le changement climatique peut avoir une incidence importante sur la composition génétique et la structure des communautés, et peut modifier la connectivité génétique entre les populations, augmentant le risque de perte de la diversité génétique. Le changement climatique influe normalement de façon disproportionnée sur certaines espèces de communautés, réduisant la synchronie et la symétrie entre espèces en interaction telles que les prédateurs et les proies. Le changement climatique peut aussi agir de façon synergique avec d’autres processus, par exemple la fragmentation de l’habitat, la maladie et les espèces envahissantes, pour exacerber les effets globaux. Comme les réactions des diverses espèces au changement climatique varient, certaines s’adapteront et resteront au sein d’une communauté, tandis que d’autres la quitteront et que des espèces non indigènes pourront s’y intégrer. La conséquence est l’apparition potentielle de nouvelles communautés biotiques, ce qu’on appelle le réassemblage de la communauté. La composition de la communauté se trouve ainsi modifiée, ce qui est susceptible d’amener des changements dans la biodiversité, les interactions entre espèces, la structure trophique et les processus écosystémiques. Dans le présent rapport, nous discutons des effets potentiels au niveau de la communauté du changement climatique sur les écosystèmes terrestres, mettant l’accent sur la faune, et nous déterminons les lacunes dans les connaissances. Nous faisons également des recommandations relativement à la gestion, aux besoins en matière de recherche et aux stratégies d’adaptation. ii

Acknowledgements

Funding for this project was provided by OMNR’s Climate Change Program and by Wildlife Research and Monitoring Section.

We are grateful to the following individuals who reviewed all or part of this document: Carrie Sadowski and Paul Gray. We thank Trudy Vaittinen for report layout and production. CLIMATE CHANGE RESEARCH REPORT CCRR-36 viii

Contents Summary...... i Résumé...... i Acknowledgements...... ii 1.0 Introduction...... 1 2.0 Genetic change...... 4 2.1 Adaptation...... 4 2.2 Population size and inbreeding...... 5 2.3 Hybridization...... 6 3.0 Synergy...... 7 3.1 Habitat loss and fragmentation...... 7 3.2 Pathogens and parasites...... 8 3.3 Invasive species...... 11 4.0 Asynchrony and asymmetry...... 12 5.0 Community reassembly ...... 16 5.1 Breakdown of co-evolved interactions...... 18 5.2 Uncertainty...... 18 5.3 Resilience...... 19 5.4 Regime shifts...... 19 6.0 Recommendations...... 20 7.0 Conclusions...... 22 References...... 24 Appendix 1. Glossary...... 35 Appendix 2. Summary of studies...... 36 vi CLIMATE CHANGE RESEARCH REPORT CCRR-36 CLIMATE CHANGE RESEARCH REPORT CCRR-36 1

1.0 Introduction

Climate represents the general weather conditions of a region, including temperature, precipitation, humidity, wind, and other variables, over a long period of time (Garbrecht and Piechota 2006). Climate is affected by interactions between the atmosphere, the ocean, the land surface, the biosphere, and sea ice, as well as latitude, movements of wind belts, topography, and other variables (IPCC 2007). While natural variability in the earth’s climate has always existed, over the last century human activities have dramatically increased the rate and degree of climate change (Houghton et al. 2001, IPCC 2007). One of the key causes of current global warming are elevated levels of greenhouse gases, which are the highest they have been for the last 420,000 years (Petit et al. 1999, Houghton et al. 2001).

General circulation models of the earth’s climate project that during this century global temperatures may increase by 1.1 to 6.4 °C (IPCC 2007). Mean global surface temperature has already increased by approximately 0.74 °C since the late 1800s (IPCC 2007). Some of the projected changes include global surface temperature increases, precipitation changes (rain, snow, and ice), increased intensity of extreme weather events, sea level rise, reduced snow cover, and reduced sea ice (Galley et al. 2004). Areas at high latitudes, such as Ontario, are projected to be affected more than those at lower latitudes, such as the tropics (IPCC 2007). For example, the projected annual mean temperature increase for Canada’s terrestrial ecosystems is 3.1 to 10.6 °C by the 2080s, which is almost double the projected global average temperature change (IPCC 2006). Between 1948 and 2008, average temperatures in Ontario increased by up to 1.4 °C, but changes were more pronounced in the boreal forest and Hudson Bay lowlands regions (Environment Canada 2013). By the end of the century, the average annual temperature in the province is projected to rise by approximately 5 °C (Figure 1; Colombo et al. 2007).

Figure 1. Projected change in average annual temperature in Ontario for 2071 to 2100 compared to the 1971 – 2000 period, using version 2 of the Canadian Coupled Global Climate Model (CCGCM-A2) (Colombo et al. 2007). 2 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Climate change is a major threat to biodiversity and an important influence on species interactions and ecosystem function. There is ample evidence that ecological responses to contemporary climate change are already occurring. In the Northern Hemisphere, many taxa show a consistent trend of northward or westward expansion of their ranges as well as altitudinal shifts (Thomas et al. 2001, Walther et al. 2002, Parmesan and Yohe 2003, Walther 2010). In Ontario, range expansions may increase biodiversity due to the introduction of new species in southern regions (Kerr and Packer 1998); however, range contractions and species loss are also likely to be prevalent across northern regions. Over the next century, the climate envelope of species may shift as much as 300 to 700 km north (Rizzo and Wiken 1992, McKenney et al. 2007). For example, the extent of the boreal forest bioclimatic envelope could be reduced by as much as 50%, with more southern areas being replaced by temperate bioclimatic envelopes (Rizzo and Wiken 1992, Malcolm et al. 2002, Gray 2005). Globally, rising temperatures have also caused the advancement of spring phenology (Root et al. 2003, Edwards and Richardson 2004, Parmesan 2006). As well, the introduction of southern competitors and pathogens (such as the Virginia opossum, Didelphis virginiana, and raccoon roundworm, Baylisascaris procyonis), increased risk of cold-adapted species (such as the Canada lynx, Lynx canadensis, and American marten, Martes americana), and selection for early breeding (e.g., frog communities and muskrat, Ondatra zibethicus) have been noted (Pounds et al. 2006, Post and Forchhammer 2008, van der Wal et al. 2008, Bowman and Sadowski 2012). These changes appear to be systematic trends with considerable long- term consequences. In fact, it has been suggested that the effects of climate change on biodiversity will likely exceed the negative effects of habitat loss due to factors other than climate change such as urbanization (Sala et al. 2000, Thomas et al. 2004, Jetz et al. 2007).

Documentation of the effects of climate change in Ontario at the species level (e.g., range shifts) is progressing; however, extrapolating climate change research from populations to communities and ecosystems is difficult (Kareiva et al. 1993, Schmitz et al. 2003, Varrin et al. 2007, Tylianakis et al. 2008, Berg et al. 2010, Fenton and Spencer 2010). Climate change can amplify the effects of other major extinction drivers, such as habitat loss, disease, and invasive species. As well, species responses to climate change are connected through simultaneous interactions with other species or adjacent trophic levels (Harrington et al. 1999, Tylianakis et al. 2008, Van der Putten et al. 2010), and temporal and spatial overlap affect biotic interactions, both of which are highly influenced by climate variables (Walther et al. 2002). As such, complex networks of biotic interactions may be disrupted (Mora et al. 2007; Brooke et al. 2008), and synchrony in ecological systems (e.g., the lynx–hare cycle) may be reduced (Stenseth et al. 2002). However, the ability to anticipate biotic responses to climate change is limited to some degree by uncertainty about how species will respond, as well as how local climates will be affected by the complex, interactive effects of global changes (Houghton et al. 2001, Humphries et al. 2004, IPCC 2007). As such, predicting the effects of climate change on communities and species interactions is a challenge.

The single species effects of climate change were recently documented in the climate change research report entitled The Known and Potential Effects of Climate Change on Biodiversity in Ontario’s Terrestrial Ecosystems (Varrin et al. 2007). Three general mechanisms that can affect species as a result of climate change were identified: demographic, phenological, and genetic, which can each result in either population expansions or contractions, depending on the ecology of particular species (Varrin et al. 2007). In addition to the species-specific effects of climate change, the potential effects of climate change on terrestrial communities remain of great concern. For example, long-standing species interactions and ecosystem services may be disrupted. As such, in this update of the review by Varrin et al. (2007), we have chosen to focus on the second part of that report, i.e., biotic interactions and the potential effects of climate change on terrestrial communities, as this is where the greatest uncertainty remains. Varrin et al. (2007) proposed four categories of climate change effects on biotic interactions: asymmetries, asynchronies, synergies, and thresholds. In this report, we elaborate on these topics with discussions of synergies, asynchrony, and asymmetry, and the outcome of these processes, i.e., community reassembly (Figure 2). We have omitted the threshold category as we believe thresholds can occur in all categories of community-level climate change effects (e.g., Folke et al. 2004). We begin with a brief review of the effects of climate on genetic change because we felt that recent research was sufficient to warrant an update of the information on this topic provided by Varrin et al. (2007). CLIMATE CHANGE RESEARCH REPORT CCRR-36 3

Figure 2. A schematic depiction of the potential effects of climate change on communities. Classes of effects are synergy, asymmetry, and asynchrony, all of which can potentially culminate in community reassembly.

We first summarize the potential genetic effects of climate change on terrestrial populations and communities, with a focus on wildlife. Second, we report on synergies between climate change and other extinction drivers, such as habitat fragmentation. Third, we discuss asynchronies and asymmetries between interacting species. And lastly, we discuss community reassembly, the outcome of these community-level climate change effects, and its resulting effects on species interactions. We also make recommendations for associated management considerations, research needs, and response strategies. In addition, we have included a glossary of technical terms (Appendix 1) used in this document, and an updated review of climate change studies of vertebrate species that occur in Ontario (Appendix 2). We updated the review by Varrin et al. (2007) by evaluating studies, including peer-reviewed journal articles or books published since 2006 inclusive, in which long-term data (>5 years) were quantitatively assessed for population responses to changing climate. Our review combined with that by Varrin et al. (2007) indicated that, overall, the longer-term effects of climate change have been studied on 181 species that occur in Ontario. Of these species, effects are reported as equivocal for 101, consistent with range expansion for 68, and consistent with range contraction for 12. 4 CLIMATE CHANGE RESEARCH REPORT CCRR-36

2.0 Genetic change

Summary 2.1 Adaptation Climate change can initiate range When faced with new selection pressures caused by a changing climate, species can disperse to suitable habitats elsewhere, adapt via expansions and contractions, changes phenotypic plasticity (without change in genotypes), adapt via genetic in individual breeding behaviour, and change (i.e., microevolution, a genetic response to consistent selection population , all of which may on heritable traits), or face extinction (Holt 1990,Visser 2008, Nicotra et significantly affect the genetic composition al. 2010, Chen et al. 2011, Hoffmann and Sgro 2011). and structure of species, populations, Phenotypic plasticity, the ability of individuals to modify their behaviour, morphology, or physiology in response to altered and communities. The rapid northward environmental conditions, allows individuals to adapt to a rapidly expansion of some species may lead to changing environment (Walter et al. 2002, Price et al. 2003, Yeh and increased secondary or renewed contact Price 2004). Phenotypic responses to climate change can include between species and populations resulting changes in behaviour (Visser et al. 2004, Jonzen et al. 2006, Both in increased incidence of hybridization. 2007), distribution (Parmesan 2006, Pounds et al. 2006, Hitch and Leberg 2007), and morphology (Yom-Tov 2001). Phenotypic changes This may negatively affect species allow organisms to cope with short-term environmental change; through loss of diversity and fitness however, microevolution, which involves genetic modifications, is declines, but positive effects are also thought to be essential for the persistence of populations faced with possible, as in the hybrid vigour noted in long-term directional changes in the environment (Lande and Shannon 1996). Evidence indicates that such microevolutionary adaptation recolonizing populations of fishers (Martes has occurred in several species in response to contemporary pennanti) in Ontario. Evidence also climate change. Réale et al. (2003) demonstrated that red squirrels indicates that climate change can alter (Tamiasciurus hudsonicus) in western Canada advanced breeding by genetic connectivity among populations 18 days over 10 years in response to warmer spring temperatures and and many populations are predicted to increased spruce cone abundance. Part of this phenological change resulted from phenotypic plasticity (87%), but a smaller proportion of this decrease in size as a consequence of shift resulted from genetic changes (13%), potentially representing a climate change, increasing the risk of rapid evolutionary response to selective pressures resulting from climate losing genetic variation due to genetic drift. change (Réale et al. 2003, Berteaux et al. 2004). As well, evolution towards greater dispersal has been documented in several species of When faced with new selection insects. In the United Kingdom, two species of wing-dimorphic bush pressures caused by changing climate, crickets (Metrioptera roeselii, Conocephalus discolor) have evolved species can disperse to suitable habitats longer wings at their northern range boundary, with mostly long-winged elsewhere, accommodate the changes forms participating in a range expansion, while short-winged forms did not move farther north (Thomas et al. 2001).The relative influence of via phenotypic plasticity, adapt via both plasticity and evolutionary adaptation on population persistence genetic change, or face extinction. In the in a changing environment will likely depend on species characteristics short-term, phenotypic change is likely such as generation time, mating system, dispersal capacity, the strength to be a more important mechanism for and direction of selection, and the presence of ecologically relevant coping with changing environmental genetic variation (Anderson et al. 2012). Overall though, ecological plasticity is likely to be more important than evolutionary change as a conditions than evolutionary change; mechanism to cope with changing environmental conditions in the short- however, as climate change accelerates, term, as plasticity acts within a generation, whereas evolutionary genetic plastic responses may be inadequate changes involve multiple generations (Williams et al. 2008). However, for providing long-term solutions to the there are limits to the extent of plastic responses, and they may be inadequate for providing long-term solutions to the challenges faced by challenges to species survival. species as climate change accelerates (Figure 3) (DeWitt et al. 1998, de Jong 2005). CLIMATE CHANGE RESEARCH REPORT CCRR-36 5

Fast Poor response to short- term changes, good Good response to changes response to long-term  stable fitness changes  fitness decreases

Good response to short- Poor response to short and term changes, poor long term changes response to  fitness decreases long-term changes

 fitness decreases Contemporary evolution Slow Low High Phenotypic plasticity

Figure 3. Qualitative predictions of the response of a population to rapid environmental change (such as current and predicted climate change), based on the level of phenotypic plasticity and rate of contemporary evolution in the population (redrawn with permission after Berteaux et al. 2004).

2.2 Population size and inbreeding Evidence indicates that climate change can alter genetic connectivity among populations and as a result many populations are predicted to decrease in size (Møller et al. 2004). Smaller population sizes and reduced gene flow will most likely lower effective population size, and thereby increase the risk of losing genetic variation due to genetic drift (Frankham 1999, Cobben et al. 2012).

In Yosemite National Park, USA, changes in genetic diversity for populations of two species of small mammals have been observed to differ in response to climate warming (Rubidge et al. 2011). The alpine chipmunk (Tamius alpinus) has retracted its elevational range upwards as a result of a 3 °C temperature increase over the last 100 years. Conversely, the closely related and ecologically similar lodgepole chipmunk (T. speciosus) maintained a stable elevational range over the same period. Between the two time periods, T. alpinus showed increased genetic subdivision and loss of overall genetic diversity, with a significant decline in average allelic richness. As well, only modern T. alpinus populations showed significant isolation by distance. In contrast, T. speciosus showed no significant changes in population structure, overall gene diversity, or richness. These results strongly support a climate-driven range contraction that has resulted in a loss of genetic diversity and increased local isolation for alpine chipmunk populations (Rubidge et al. 2011). As the climate continues to warm, these and other montane species are likely to further contract their elevational range, experiencing further losses of genetic diversity and population fragmentation (Epps et al. 2006, Moritz et al. 2008). Genetic diversity is important for mitigating climate change effects, and loss of genetic diversity may signal demographic collapse and reduced fitness (Spielman et al. 2004, Hoffmann and Sgro 2011).

Similar processes appear to be underway in Ontario. Since the 1970s, Canada lynx populations have contracted at their southern range edge by almost 200 km and current populations along the contracting edge exhibit lower genetic variability than core lynx populations. The proximate cause of reduced genetic variability at range edges appears to be warm winter temperatures, although changes in forest composition may also play a role (Koen et al. 2014. Small population sizes will also lead to increased risk of inbreeding and inbreeding depression (Rowley et al. 1993, Kruuk et al. 2002). A long-term study of red-cockaded woodpeckers (Picoides borealis) found that inbred females are not adjusting their egg-laying date as the climate warms and, as such, their time of breeding no longer coincides with optimal foraging conditions for prey, such as insect larvae (Schiegg et al. 2002). However, females that are not inbred are laying eggs earlier than before, exhibiting phenotypic plasticity. By unequally affecting inbred and non-inbred individuals, climate change may pose an additional threat to (Azevedo et al. 2000). 6 CLIMATE CHANGE RESEARCH REPORT CCRR-36

2.3 Hybridization Global climate change can shift climate regimes, leading to species range shifts and possibly increased secondary contact between recently diverged species (Parmesan 2006). For example, during a series of warm winters between 1995 and 2003, the southern flying squirrel (Glaucomys volans), a specialist of eastern temperate deciduous forests, rapidly expanded its northern range limit by approximately 200 km (Bowman et al. 2005). The range expansion brought G. volans into increased sympatry with its boreal forest counterpart, the northern flying squirrel (G. sabrinus), and resulted in the formation of a new hybrid zone in central Ontario (Bowman et al. 2005, Garroway et al. 2010).

In Canada’s western Arctic, grizzly bears (Ursus arctos) have been increasingly present in polar bear (Ursus maritimus) territory (Kelly et al. 2010). Wild polar–grizzly hybrids and second-generation offspring have been documented in the northern Beaufort Sea of Arctic Canada (Miller et al. 2012). As the climate continues to warm, polar bears will likely be forced to spend increasingly more time on land due to the melting of the polar ice caps and shorter seasons of sea ice cover, perhaps even during the breeding season, bringing them into closer contact with grizzly bears (Miller et al. 2012). Similarly, interbreeding among other Arctic species could significantly affect polar biodiversity. For example, hybridization between the endangered North Pacific right whale and the more numerous bowhead whale could quickly push the former to extinction (Kelly et al. 2010). Lynx × bobcat (Lynx rufus) hybrids may occur in Ontario as the bobcat expands its range north, however these hybrids are expected to be relatively rare due to the relatively old divergence of this species pair; competition may be a more important process than hybridization in determining the effect of climate change on these two species (Bowman and Sadowski 2012).

Hybridization can be detrimental to species because of diversity loss, and fitness declines following admixture (Rhymer and Simberloff 1996, Muhlfeld et al. 2009). However, hybridization is one of the few mechanisms leading to new combinations of genes, which can facilitate evolutionary adaptation by introducing genetic variation (Hoffmann and Sgro 2011). For example, interspecies hybridization in Darwin’s finches has introduced the genetic variance in morphology needed for adapting to changing climate conditions (Grant and Grant 2010). Meanwhile, in Ontario, fishers appear to exhibit hybrid vigour between recolonizing populations (Carr et al. 2007b). Therefore, as species range shifts occur and the incidence of hybridization increases, there may be unexpected evolutionary consequences and even benefits, such as improving adaptive capacity, when new variation is introduced into populations (Hoffmann and Sgro 2011). CLIMATE CHANGE RESEARCH REPORT CCRR-36 7

3.0 Synergy

Summary 3.1 Habitat loss and fragmentation Habitat loss and fragmentation are two of the primary drivers of contemporary A synergy is an interaction of species extinctions (Mainka and Howard 2010). When habitat loss occurs, processes such that the total populations are at increased risk of extinction (Bender et al. 1994, Fahrig effect is greater than each 2001). Furthermore, habitat fragmentation increases isolation between habitats, process acting independently. reducing population connectivity (Opdam 1991, Debinski and Holt 2000). Lack of connectivity, in turn, leads to reduced recolonization of locally extinct habitat The synergistic effects of patches, further increasing the probability of extinction over time across the habitat fragmentation, habitat whole landscape or metapopulation (Brown and Kodric-Brown 1977, Hanski and loss, and climate change Gilpin 1991). Significant changes in species’ populations and distributions have are expected to contribute already been detected in response to the effects of each of these processes to the decline of biological acting independently (Fahrig 2003). However, growing evidence suggests that the synergistic effects of habitat fragmentation, habitat loss, and climate change diversity. Populations in will also contribute significantly to the decline of biological diversity (Opdam and fragmented landscapes Wascher 2004, McLaughlin et al. 2005, Brooke et al. 2008), and the potential are more susceptible to combined effects of these processes may be greater than those estimated environmental stressors, individually (de Chazal and Rounsevell 2009). such as climate change, Populations in fragmented landscapes are more susceptible to environmental than those in connected stressors, such as climate change, than those in continuous landscapes (Meffe landscapes. Habitat and Carroll 1997, Travis 2003). Yet climate change studies often presume fragmentation increases that other habitat features in the environment are uniform; therefore, shifts in isolation between populated species geographic range are attributed to climate, while effects of landscape composition and configuration are not accounted for (Opdam and Wascher 2004). habitats, and reduces However, the assumption of uniform habitat does not hold true for many parts of population connectivity, which Canada, where the most intensive land uses and the greatest level of landscape in turn increases the risk of fragmentation are concentrated in biodiversity hotspots, such as southern Ontario extinction. Regions of Ontario (Kerr and Cihlar 2003). In today’s anthropocentric world, areas of unsuitable with the most intensive land landscape and man-made barriers such as highways, agricultural zones, and cities may impede species’ movements. The resulting barriers to population uses and the greatest level connectivity among habitat patches will likely decrease dispersal (Wasserman et of landscape fragmentation, al. 2012), increase mortality (Fahrig et al. 1995), reduce genetic diversity (Reh such as southern Ontario and Seitz 1990, Wasserman et al. 2012), reduce recolonization following local which is also the most extinction (Semlitsch and Bodie 1998), and may ultimately lead to population biologically diverse area of declines (Brown and Kodric-Brown 1977). For example, a rapid population decline the province, are particularly of the green salamander (Aneides aeneus) within a highly fragmented habitat in the southern Appalachians, USA, has been linked to an increase in temperatures at risk. Similar synergies over the last 50 years (Corser 2001). As well, it is predicted that by the year 2100 may occur between climate as many as 1800 of the world’s land bird species could be threatened by the change and pathogens, synergistic effects of climate change and land conversion (Jetz et al. 2007). whereby climate change Species’ distributions are limited by bioenergetic constraints, suggesting that facilitates the spread and global warming will allow many species to expand northwards (Humphries et al. effect of novel pathogens, 2002). Theoretically, population expansion should be fastest in regions where and between climate change landscape structure enhances dispersal, and should lag behind in regions where and invasive species. landscapes are fragmented. Warren et al. (2001) found that a butterfly range expansion in the United Kingdom did not occur in heavily fragmented landscapes. In spite of the improved habitat availability caused by climate warming, 93% of the butterfly species with small dispersal capacities declined, while most of the 8 CLIMATE CHANGE RESEARCH REPORT CCRR-36

species that did expand their ranges had large dispersal capacity. They concluded that the negative effect of habitat fragmentation on species distribution was overshadowed by the positive effect of a warmer climate. In general, we should expect asymmetric selection for species with good dispersal ability over those with poor dispersal ability (Kotiaho et al. 2005), and the synergy between habitat loss and habitat fragmentation will likely magnify this effect. In Ontario, the northward range expansion of both the hooded warbler (Wilsonia citrina) and the southern flying squirrel appears to have been simultaneously facilitated by climate warming and limited by habitat fragmentation (Bowman et al. 2005, Melles et al. 2011).

Habitat fragmentation can be caused by natural disturbances (Opdam and Wiens 2002). Some species have adapted to unpredictable habitat availability by developing high mobility, and consequently are less susceptible to human-induced fragmentation. These include species from coastal habitats and early successional stages of ecosystems as well as the boreal forests of Ontario, where many species have adapted to fire disturbance. Conversely, species in systems with relatively stable natural dynamics, such as tropical rain forests, have evolved under fairly predictable conditions in a more or less continuous habitat and are therefore likely to be more susceptible to fragmentation (Opdam and Wascher 2004). Moreover, the effect of fragmentation will vary among ecosystem types. Some have argued that fragmentation effects should be strongest at high levels of habitat loss (Fahrig 1997, Swift and Hannon 2010). Forests, grasslands, and wetlands often become highly fragmented with habitat loss, whereas shrublands, farmland, and pastures are regarded as less vulnerable (Mantyka-Pringle et al. 2012). As such, forests, grasslands, and wetlands, and the species that occur within them, are likely to be vulnerable to the synergistic effects of habitat conversion and climate change. Some positive effects of the interaction between habitat fragmentation and climate change may also occur. For example, higher temperatures might result in areas that were unsuitable for colonization by certain plants to become habitable, resulting in patches added to the habitat network and the overall improvement of the spatial cohesion of some landscapes (Thomas et al. 1999).

As time progresses, landscapes dominated by human land use, such as southern Ontario, will likely continue to change due to increasing urbanization, agricultural development, and economic activity, causing further habitat fragmentation. In landscapes most vulnerable to the synergistic effects of climate change and fragmentation, the development of ecological connectivity zones, networks of narrow corridors, and wildlife passages may help to lessen the negative effects on some species (Wasserman et al. 2012).

Case study: Marten The American marten is associated with extensive snow pack, older forests, and the distribution of a competitor, the fisher (Carroll 2007, Krohn et al. 1995). Snow allows the marten, with its small ratio of body mass to foot area, to gain a competitive advantage over sympatric carnivores and may also affect prey abundance and vulnerability (Krohn et al. 1995). Climate change is projected to result in increases in winter temperature in many areas, which is likely to result in a decrease in winter snowpack and migration of forest communities upward in latitude and elevation (IPCC 2007, Littell et al. 2011). All of these changes will disadvantage the marten. Marten also have large area requirements, and thus are expected to be vulnerable to landscape change (Cardillo et al. 2006).

As such, climate change and its synergistic effects with habitat fragmentation are likely to affect American marten populations. Carroll (2007) examined these combined effects for marten in southeastern Canada and the northeastern United States, and found that marten populations showed stronger declines due to climate change alone than due to overharvest or logging, but climate change interacted with logging (which results in habitat loss and fragmentation) to increase overall vulnerability. This highlights the potential threats faced by small and semi-isolated populations, as climate change can interact with habitat conversion to form an (Carroll 2007, Gilpin and Soulé 1986). CLIMATE CHANGE RESEARCH REPORT CCRR-36 9

3.2 Pathogens and parasites Climate change can play a role in altering the dynamics and ecology of wildlife disease. Pathogens and their vectors are sensitive to changes in temperature, rainfall, and humidity (Harvell et al. 2002), thus climate warming can affect the distribution, seasonality, and severity of diseases (Le Conte and Navajas 2008).

Most pathogens and vectors, such as insects, have limited temperature and humidity ranges for survival and optimal reproduction. Indeed, many are limited by cold temperatures. Warmer temperatures could increase the incidence of disease both by increasing the vector population size and distribution, and by increasing the length of time vectors are present in the environment. If global temperatures, precipitation, and humidity rise, as is projected by climate change models (IPCC 2007), pathogens and vectors that are normally restricted to warmer, wetter, and lower altitude zones will be able to expand their range to previously inhospitable latitudes and altitudes leading to the exposure of naïve host populations (Kaeslin et al. 2012).

Vector-borne diseases have been predicted to increase at higher latitudes and altitudes under warming temperatures (Kuhn et al. 2005, Ogden et al. 2006). Lyme disease, a bacteria spread by some species of ticks, is currently uncommon in Canada, where established populations of vectors are limited to southern Ontario, Nova Scotia, and British Columbia. However, models suggest that the geographic range of tick species that transmit Lyme disease may expand significantly due to climate change, with a northern expansion of about 200 km projected by the year 2020 (Figure 4; Ogden et al. 2006). This expansion would likely be due to longer growing seasons resulting from warmer temperatures and decreased tick mortality during milder winters (Lindgren and Gustafson 2001). Seasonal tick activity under climate change scenarios suggests endemic cycles of Borrelia burgdorferi, the causative agent of Lyme disease, will be maintained in newly established tick populations (Ogden et al. 2006). As well, transmission of the bacterium to humans is often increased when warmer temperatures in the early spring result in the overlap of feeding activity of nymphal (virus infected) and larval (uninfected) Ixodes scapularis ticks. Under these weather conditions, infection is more readily passed from infected ticks to uninfected ticks through small rodents. Because the viral infection is brief in tick-infested rodents, feeding of both stages of tick at the same time results in more infected larval ticks and greater risk for Lyme disease infection in humans (Gatewood et al. 2009). In North America, tick-borne diseases such as babesiosis, anaplasmoses, and Powassan encephalitis, as well as mosquito-borne diseases such as dengue and West Nile virus, may also expand their ranges if there is a northern expansion of vector populations (Epstein 2001, Greer et al. 2008).

Figure 4. Projected upper temperature limits for Ixodes scapularis establishment in Canada. The graph shows the current upper geographic limits and projected limits for the 2020s, 2050s, and 2080s, assuming continuous population growth, regionally oriented economic development, and no reduction in greenhouse gas emissions. Modified, with permission, from Elsevier (Ogden et al. 2006 and Greer et al. 2008). 10 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Climate change is expected to increase the frequency of extreme weather events that affect disease cycles (de la Rocque et al. 2008). For example, in Africa, outbreaks of Rift Valley fever, a mosquito-borne disease, have been linked with incidences of higher seasonal rainfall. Many insect vectors have population booms associated with large amounts of rain, and the flooding that accompanies heavy rainfall can increase the spread of waterborne pathogens. Conversely, decreased rainfall and drought can result in animals congregating around limited food and water resources, thereby increasing population densities and possibly increasing pathogen and parasite transmission (Kaeslin et al. 2012).

Climate change may also affect the immune status of host animals due to heat or nutritional stress. If increased temperatures or extreme weather events limit the availability or abundance of food, animals may become more susceptible to heavy parasite loads and may experience increased exposure and susceptibility to pathogens (Kaeslin et al. 2012). For example, survival of the brain worm (Parelaphostrongylus tenuis) of white-tailed deer (Odocoileus virginianus) may have increased due to recent warmer temperatures and milder winters in the northcentral United States and southern Canada. The parasite, which overwinters as larvae in snails, causes neurological disease in moose (Alces alces) and caribou (Rangifer tarandus). Moose are already experiencing health repercussions (such as increased heart rate and weight loss) due to heat stress caused by recent climate warming (Lenarz et al. 2009), and may therefore be more at risk of contracting parasitic and infectious diseases (Murray et al. 2009). Similarly, amphibians suffering from climate change induced stresses, such as increased ultraviolet radiation, may be more susceptible to pathogens (Harvell et al. 2002).

Due to climate warming, southern species such as the grizzly bear, red fox (Vulpes vulpes), and white-tailed deer have shifted their ranges north towards the Arctic (Kaeslin et al. 2012). These southern species bring diseases for which their Arctic counterparts, such as polar bear, Arctic fox (Vulpes lagopus), and caribou, have no immunity. For example, brucellosis, a bacterial disease found in cattle, dogs, wild animals, and humans, has now been found in baleen whales (Mysticeti spp.) (Kaeslin et al. 2012). Meanwhile, since 1995, the geographic range of the lung parasite (Parelaphostrongylus odocoilei) of caribou has shifted northward from the Pacific coastal range of the United States to include Alaska, and from British Columbia, Canada, to include the Yukon and Northwest Territories (Hoberg et al. 2008). Warmer summer temperatures also now allow lung nematode (Umingmakstrongylus pallikuukensis) larvae, often found in muskoxen, to develop to the infectious stage within the intermediate host, the marsh slug (Deroceras laeve), at a rate that has reduced the parasite’s life cycle from 2 years to 1 year (Kutz et al. 2005). This means that muskoxen are now exposed to an increased intensity of infection and are infected earlier in the season and at younger ages. The parasite can compromise the respiratory system, and thus can have adverse effects on muskoxen fecundity, predation rates, and survival (Kutz et al. 2001).

Climate-driven changes in habitat and resources may also force animals to shift their ranges or to alter their migration routes into new ecosystems where they may introduce or be exposed to novel pathogens (Kaeslin et al. 2012). Conversely, climate warming could make environmental conditions on breeding grounds more favourable for year-round survival, replacing migratory populations with year-round resident populations (Lusseau et al. 2004, Bradshaw and Holzapfel 2007). Migrations can be beneficial by allowing hosts to escape the continual build-up of pathogens in the environment (Loehle 1995, Altizer et al. 2003) or by eliminating infected animals from the population during arduous migrations (Gylfe et al. 2000, Bradley and Altizer 2005). Altered migration routes and range shifts could result in migratory animals encountering and transferring pathogens to previously naïve host populations, or themselves becoming exposed to novel infectious diseases. Pathogens introduced into previously unexposed host populations can spread quickly, cause high fatality rates, and lead to significant host population declines (Harvell et al. 2009).

The climate is changing at an unprecedented rate, altering physical and biological processes, including patterns of infectious disease. Climate change is expected to increase levels of infection, change the distribution of diseases and parasites, affect host population dynamics, and have cascading ecological, sociological, and economic effects. As well, changes in the distribution and abundance of diseases and parasites will have significant implications for natural resource agency programs and the public at large. As such, research and monitoring of wildlife diseases should be encouraged, so that both natural resource and public health agencies have time to prepare response strategies when diseases begin to spread into new areas. CLIMATE CHANGE RESEARCH REPORT CCRR-36 11

3.3 Invasive species Biological invasions occur when a species is introduced to a habitat or ecosystem where it is not native and subsequently becomes established. Invasive species can reduce biodiversity and alter the structure and function of entire ecosystems (MacDougall and Turkington 2005, Mainka and Howard 2010, Vila et al. 2010). As a result of these effects, biological invasions are an important threat to biodiversity and ecosystem services, and are considered one of the five largest threats to ecosystem integrity (MEA 2005).

Recent research suggests that climate change is likely to interact with and affect the distribution, spread, abundance, and effects of invasive species (Gritti et al. 2006). Climate change may influence invasive species and their effects on species, populations, and ecosystems in several ways. First, global warming could provide new opportunities for introductions to areas where, until recently, those species were not able to survive. Species introduced from warmer regions to temperate areas have, until recently, been constrained by growing seasons that were too short or winter temperatures that were too cold, which prevented them from becoming naturalized (Walther et al. 2009). With warmer temperatures, some species may be able to extend their reproductive period and expand their northern range limits (Walther et al. 2002). For example, a strong association between patterns of the emergence of the invasive gypsy moth (Lymantria dispar) and climatic suitability is evident in Ontario (Régnière et al. 2009). Records indicate a significant increase in the distribution of the invasive moth since 1980 during which time the climate has warmed. However, between 1992 and 1997, a temporary decline in climatic suitability occurred and resulted in a drastic reduction in the area defoliated by these moths. Since 1998, the warming trend has continued, and resultant defoliation is expected to threaten hardwood forest resources as climate change allows the gypsy moth to expand farther north and west. It is estimated that by 2050 the proportion of Canada’s deciduous forests at risk of gypsy moth damage will grow from the current 15% to more than 75% (Régnière et al. 2009).

In addition to the removal of physiological constraints, climate change can also affect dispersal of species in various ways. For example, warmer nocturnal temperatures increase flight activity of invasive winter pine processionary moth (Thaumetopoea pityocampa) females, enabling them to disperse over greater distances (Battisti et al. 2006). As well, increasing temperatures could result in an additional generation of the invasive moth each year (Walther et al. 2002). Meanwhile, the historic range of the North American native mountain pine beetle (Dendroctonus ponderosae) has been limited by climate. However, as a result of increased warming at higher latitudes and altitudes, the beetle is able to complete a life cycle in one season rather than the typical two, allowing for more rapid range expansion into new environments (Logan and Powell 2001).

Invasive species can have major effects on the communities and ecosystems they invade, where they may dominate function or richness and transform ecosystem properties, which inevitably leads to changes in biological communities (Richardson et al. 2000, Vila et al. 2009). By definition, invasive species are typically successful and abundant, whereas many native species are rare and constrained. Invasive species also tend to have characteristics that differ from non-invasive species, which may provide them with a competitive advantage under warming climatic conditions, and allow them to take over empty niches, or compromise native species’ ability to compete against hardy generalist invaders (Mainka and Howard 2010). For example, many invasive plants have broad climatic tolerances and large geographic ranges, and also often have characteristics that facilitate rapid range shifts, such as low seed mass and short time to maturity (Rejmánek and Richardson 1996, Qian and Ricklefs 2006). Therefore, as the local environment changes, resident species may become increasingly poorly adapted, which will provide opportunities for newcomers that are better adapted and, thus, more competitive under the new conditions. For example, milder winters in central Europe changed the habitat of deciduous forests to conditions that are now more suitable for evergreen broad-leaved species (Berger et al. 2007). Acting together, climate change and invasive species can compromise the ability of many native species to survive, leading to reduced diversity of native species (Mainka and Howard 2010). These changes may subsequently alter existing species interactions, which may lead to unexpected effects on ecosystems (Tylianakis et al. 2008).

Finally, climate change may also challenge the definition of invasive species because in some areas species that were previously invasive may diminish in prevalence or effect. Meanwhile, native species may increase in abundance, and colonize new habitats taking on characteristics of exotic invaders (Hellman et al. 2008). 12 CLIMATE CHANGE RESEARCH REPORT CCRR-36

4. Asynchrony and asymmetry

Summary Variation in species’ responses to climate change can alter existing relationships, resulting in asynchrony (or a mismatch) in Rapid climate change may reduce predator–prey interactions, insect–plant interactions, migrations, synchrony in co-evolved systems and may reproduction, and phenology. For example, in recent years the have asymmetric effects, which depend on strong trophic interaction between winter moth (Operophtera brumata) egg hatching and English oak (Quercus robur) bud burst species traits. Already, the phenology and has begun to break down due to warming temperatures (Visser and distribution of many plant and animal species Holleman 2001). In warm springs, winter moth eggs were predicted have changed, from the level of individuals to to hatch up to three weeks before oak buds burst. However, newly communities and across multiple trophic levels. hatched caterpillars can only survive for a maximum of 10 days The timing of events such as leaf unfolding, without food (Wint 1983), therefore asynchrony in this relationship flowering, emergence of nymphs, arrival of can lead to increased mortality in winter moths (Visser and Holleman 2001). migratory birds and butterflies, and breeding has advanced, whereas other events such as Climate change could also alter the timing of predation events leaf fall have become delayed, leading to an (e.g., prey and predator encounters), which could result in stronger or weaker trophic interactions between predators and prey extended growing season. Estimates are that (Mølleret al. 2010). For example in Britain, newts (Triturus spp.) 62% of species, most of which occur in the have advanced the timing of their entry into ponds, whereas their Northern Hemisphere, have already shifted prey, the common frog (Rana temporaria), have not substantially their timing of spring events in response to altered their reproductive phenology (Beebee 1995). Therefore, recent climate warming. Variation in species’ embryos and larvae of early breeding frogs are now exposed to responses to climate change can alter existing higher levels of newt predation (Walther et al. 2002). In Canada, the most iconic synchronous system is the cycle between Canada lynx relationships, resulting in asynchrony (or a and snowshoe hare, which is controlled, in part, by the influence mismatch) in predator-prey interactions, insect– of the NAO (North Atlantic Oscillation) (Stenseth et al. 2002). plant interactions, migrations, reproduction, As specialized hunters, Canada lynx prey almost exclusively on and phenology. For example, North American snowshoe hare, and lynx populations are closely tied to population wood warblers (Parulidae) are not advancing cycles of snowshoe hare. Synchrony between lynx and hare is in phenology in response to climate change as greatest during cold periods, and synchrony appears to break down during periods of warming (Scott and Craine 1993). Canada fast as key prey (such as the eastern spruce lynx are highly effective deep snow hunters, therefore this pattern budworm, Choristoneura fumiferana). As may be due, in part, to an increase in specialized predation during well, the iconic lynx–snowshoe hare (Lepus cold periods as a result of changes in snow depth and structure americanus) cycle may become decoupled (Stenseth et al. 2004). As well, deep snow typically excludes the as the climate warms. In general, the more lynx’s main competitors, the coyote (Canis latrans), fisher, and specialized the relationship between species bobcat, from its winter habitat (Smith 1984, Litvaitis 1992, Murray et al. 1994, Krohn et al. 1995). Less snow cover could therefore (e.g., plants and their pollinators), the more mean more competition for lynx resulting from more predation on vulnerable each of them is likely to be to hares by other carnivores. Bobcats, coyotes, and fishers, who prey the phenological effects of climate change. on a more diverse range of prey, may be better equipped to adapt Mistiming and mismatching can reduce to a changing climate than specialists such as the Canada lynx. As individual fitness and result in population such, the lynx–snowshoe hare cycle may become decoupled as the declines, increasing the risk of population climate warms (Stenseth et al. 2002, 2004). extinctions and associated loss of biodiversity, Interspecific competition is one of the major factors determining while the decoupling of predator–prey the distribution and abundance of species and thus species relationships will likely affect other trophic levels. composition at the community level as well (MacArthur and Levins 1967). In a stable environment, competition between two species over common resources should lead to niche differentiation or local CLIMATE CHANGE RESEARCH REPORT CCRR-36 13

extinction of the weaker competitor (Hardin 1960). Changes in environmental conditions can affect the competitive relationships among species. For example, migratory bird species may be at a disadvantage compared to resident bird species because changes in their wintering areas and along migration routes do not necessarily reflect those occurring in their breeding areas (Berthold et al. 1998). Both resident and migratory species may be able to adapt to changes through selection, but individuals of resident species are expected to be better able to adjust to warming spring temperatures and an advanced phenology of their food items (Ahola et al. 2007).

Species that migrate from wintering grounds to breeding areas may also be more vulnerable to the effects of climate change because they may arrive at an inappropriate time to exploit the habitat optimally, may experience higher competition with resident species, and are involved in more inter-specific interactions that may be disrupted (Berthold et al. 1998, Lemoine and Böhning-Gaese 2003). Species of birds that migrate over long distances must co-occur with their food sources (while avoiding their enemies) in the habitat in which they grow, and then, following migration to their breeding grounds, egg-hatching must be in synchrony with the food sources that they feed to their newborns (Sillett et al. 2000). Short-distance migrants may be more flexible in their response to climate change, because the circumstances on their wintering grounds will be a better predictor for the optimal arrival time on their breeding grounds (Berthold et al. 1992, Pulido et al. 1996). For example, competition for nest-holes between resident great tits (Parus major) and migratory pied flycatchers (Ficedula hypoleuca) increases when the timing of breeding onset is closer to overlapping and when the densities of tits or pied flycatchers are high. All these factors can be affected by climate change, indicating that it has great potential to affect the level of interspecific competition between these two species (Ahola et al. 2007).

In another example, Adélie (Pygoscelis adeliae), gentoo (P. papua), and chinstrap (P. antarcticus) penguins in the Western Antarctic Peninsula breed in sequence and over a period of three weeks or less (Trivelpiece et al. 1987). This staggered breeding may reduce direct foraging competition during chick rearing (Lishman 1985, Trivelpiece et al. 1987), and is an important factor for the distribution of limited nesting space, as gentoo and chinstrap penguins can out compete Adélies for available space in mixed colonies (Carlini et al. 2005, Sander et al. 2007). However, as the climate has warmed, gentoo penguins have exhibited greater plasticity in breeding phenology, which has decreased the mean interval between Adélie and gentoo breeding in warm years, increasing competition for nesting space in mixed colonies (Lynch et al. 2012). This may be one explanation for why small Adélie populations breeding in mixed colonies with gentoo penguins have been declining in recent years (Lynch et al. 2008). As such, differential responses in breeding phenology to changing temperatures represent an additional mechanism by which climate change may affect competitive interactions (Lynch et al. 2012).

Phenology refers to the timing of plant and animal life cycle events and how these are influenced by seasonal and interannual variations in climate (Walther et al. 2002). The phenology of organisms has evolved through natural selection to match their environmental conditions and to maximize the fitness of individuals (Futuyma 1998). Under normal conditions, the timing of recurring activities in the dependent species is controlled by abiotic variables (such as temperature) such that synchronization is maintained (Visser and Holleman 2001). These response mechanisms are the result of selection under the range of conditions experienced in the past (van Noordwijk and Müller 1994). However, under novel environmental conditions, synchronization between different trophic levels can break down because natural selection on species cannot always keep pace with the rate of change in environmental conditions in a rapidly warming climate (Visser and Holleman 2001). Global warming has altered the phenology and distribution of many plant and animal species, resulting in changes from the level of individuals to communities and multiple trophic levels (Walther et al. 2002, Parmesan and Yohe 2003, Root et al. 2003). The breakdown of phenological relationships will have important consequences for trophic interactions, food–web structures, predator–prey interactions, and biodiversity (Edwards and Richardson 2004). Climate warming has advanced the timing of events such as leaf unfolding (Menzel and Fabrian 1999), flowering (Fitter and Fitter 2002), emergence of nymphs (Roy and Sparks 2000), and breeding (Forchhammer et al. 1998, Dunn and Winkler 1999, Forchhammer et al. 2002), whereas other events such as leaf fall have become delayed, leading to an extended growing season for both plants and the species that feed on them (Menzel and Fabrian 1999).

Climate change has also affected the timing of avian migration (Inouye et al. 2000). If the phenology of a species is shifting at a different rate from that of the species on which it relies (i.e., for food or pollination), this will lead to 14 CLIMATE CHANGE RESEARCH REPORT CCRR-36

mistiming of its seasonal activities (Visser et al. 2004). In the Netherlands, the pied flycatcher is currently suffering a trophic mismatch with its insect prey. The timing of peak insect abundance has advanced with climate warming, however the birds are not arriving on their breeding grounds any earlier (Both and Visser 2001). As such, the birds are suffering from mistimed reproduction. Similarly, in response to increased temperatures and decreased spring snow cover, egg laying and hatching of the greater snow goose (Chen caerulescens atlantica) occurred progressively earlier over a 16-year period (Dickey et al. 2008). However, both gosling mass and size at fledging were lower and there was an overall decline in reproductive success, in part due to trophic mismatch between the hatching date of goslings and the timing of peak plant quality (Dickey et al. 2008). In the Rocky mountains, the American robin (Turdus migratorius) is now arriving 14 days earlier than it did 2 decades ago, but as there has been no advancement of the date of snow melt, the interval between the first arrival of the robins and the first date of bare ground (which correlates with food availability) has grown by 18 days over this period (Inouye et al. 2000).

Strode (2003) suggests that North American wood warblers are not advancing in phenology as fast as their key prey (such as the eastern spruce budworm, Choristoneura fumiferana) are responding to increased temperatures. The emergence of spruce budworm occurs at approximately the same time that buds flush on host trees (Candau and Fleming 2008). Earlier bud flush in some areas may facilitate spruce budworm outbreaks. As climate change progresses, frequency and duration of spruce budworm outbreaks is predicted to increase because of the positive effect of warmer winter and spring temperatures and drought on insect physiology (Greenbank 1963, Mattson and Haack 1987) and because of the possibility of reduced synchrony between the spruce budworm and its natural enemies, such as wood warblers (Fleming 2000). As such, a substantial increase in defoliation of trees is predicted for northern Ontario (Candau and Fleming 2008).

In general, the more specialized the relationship between species, the more vulnerable each of them is likely to be to the phenological effects of climate change. For example, if successful pollination of a particular plant requires a pollinator with very specific morphological characteristics (e.g., tongue length) (Corbet 2000), but that pollinator has advanced its phenology and is no longer present during peak flowering, then that plant is more vulnerable to losing these pollinator services than are species that are visited by a wide range of pollinator species. Climate change may affect co-occurrences of plant and pollinator species spatially as well as temporally. Range shifts in plants (e.g., Lenoir et al. 2008, Pompe et al. 2008, Thuiller et al. 2008) and pollinators (e.g., Parmesan 1996, Parmesan et al. 1999, Menéndez et al. 2007, Settele et al. 2008) are occurring, but overlaps in current species distribution may not persist. For example, Schweiger et al. (2008) modelled the climatic niche for the butterfly Boloria titania and its host plant Polygonum bistorta and found that the overlap of their climatic niches will be considerably reduced under projected climate change scenarios. However, while most incidences of asynchrony are expected to negatively affect the species involved, improved matching of beneficial interactions (e.g., pollination) or more mismatching of adverse interactions (e.g., release of a plant from its herbivore) may also occur (Visser and Holleman 2001).

Climate change is likely to disrupt existing species interactions by altering the temporal and spatial nature of events; however, the direction and magnitude of these shifts are difficult to predict. This difficulty arises because species and populations: (i) differ in the extent to which their life history events (such as breeding) are able to accelerate with warming, (ii) experience different warming trends due to variations in mean seasonal timing of events and microhabitat use, (iii) vary in the extent to which their phenological responses are driven/constrained by factors other than temperature, and (iv) may respond to changing climate in other ways, such as through distributional changes (Thackeray et al. 2010, Visser and Both 2005).

It has been estimated that 62% of species, most of which occur in the northern hemisphere, have already shifted their timing of spring events (such as earlier frog breeding, bird nesting, and arrival of migratory birds and butterflies) in response to recent climate warming (Parmesan and Yohe 2003), with different taxonomic groups and trophic levels showing different magnitudes of response (Parmesan 2007, Thackeray et al. 2010). As well, a significant number of species range shifts have been recorded (Parmesan and Yohe 2003, Walther et al. 2002). If species that rely on each other are indeed showing different magnitudes (or even directions) of response, then the implications may be severe for ecosystems, especially if keystone species are affected (Figure 5) (Winder and Schindler 2004, Visser and Both 2005). CLIMATE CHANGE RESEARCH REPORT CCRR-36 15

Figure 5. Distribution of two species, A and B, whose ranges largely overlap, and species’ distribution in response to climate change, where species-specific changes cause the ranges to separate. Adapted from Peters (1992).

Such mistiming and mismatching has been linked to reductions in individual fitness and population declines, increasing the risk of population extinctions and biodiversity loss (Platt et al. 2003, Winder and Schindler 2004, Both et al. 2006, Miller et al. 2008). As well, the effects of the decoupling of predator–prey relationships will likely affect other trophic levels (Winder and Schindler 2004). As such, spatial and temporal mismatches can cause drastic ecological and economic consequences due to the influence of synchrony on processes such as pollination (Elzinga et al. 2007), fisheries production (Cushing 1990), and herbivory by agricultural pests (Harrington et al. 2007).

Case study: Caribou Herbivores in the Arctic display seasonal reproduction that is timed to coincide with a peak in resource availability (Post 2003a, b). Caribou migrate between seasonal ranges and time their arrival on calving grounds to coincide with the timing of emergence of forage plants, which is crucial to the successful growth of newborn calves (Gunn and Skogland 1997). However, shifts in the timing of plant growth have already occurred at high latitudes, with plant emergence beginning earlier and lasting for a shorter period (Walther et al. 2002, Post 2003b, Forchhammer et al. 2005). As such, there is potential for a trophic mismatch between the timing of caribou arrival on their calving grounds and the timing of peak resource availability. Such a mismatch occurs when the timing of plant growth on breeding grounds advances due to warmer spring temperatures (Visser and Holleman 2001), while the timing of migration from wintering areas, which is cued by seasonal changes in day length, remains constant (Visser et al. 1998).This kind of trophic mismatch has already had negative consequences for caribou in west Greenland where temperatures have risen and forage plants have advanced their growing season by as much as 14.8 days, yet caribou calving has only advanced by 1.28 to 3.82 days, resulting in increased offspring mortality and a fourfold drop in offspring production (Post and Forchhammer 2008). As temperatures continue to warm throughout the Arctic, the extent to which plant phenology will further advance is a crucial factor in the future reproductive success of caribou (Post and Forchhammer 2008). In the Canadian High Arctic, a population of the endangered Peary caribou (R. tarandus pearyi) recently experienced a catastrophic and near-total population crash associated with increasing winter snow and ice crust formation (Miller and Gunn 2003). According to climate change projections, increasing snowfall and ice crust formation will continue to occur in this area as climate change progresses, further threatening the Peary caribou herd’s future (Miller and Gunn 2003). In Ontario, climate change is expected to affect woodland caribou (R. tarandus caribou) through habitat loss (increased incidence and severity of fires), increased energy costs (as a result of summer heat and increased harassment by insects), and increased interaction with white-tailed deer (Racey 2004, Thompson and Baker 2007). As a result, woodland caribou may be restricted to a relatively small portion of northwestern Ontario (Thompson and Baker 2007). 16 CLIMATE CHANGE RESEARCH REPORT CCRR-36

5. Community reassembly

The concepts we have discussed to this point, i.e., synergy, asynchrony, Summary and asymmetry, will lead to the formation of novel ecological communities. Individual species differ This process, known as community reassembly, is already underway in in their responses to climate Ontario. Community reassembly will have important consequences for biodiversity and ecosystem functioning. change: some species will adapt, some cold-adapted species will Species assemblages are not fixed and novel interactions are a common leave communities, and some occurrence in nature (Davis 1986, Vermeij 1991). The reassembly of warm-adapted species may communities has occurred frequently in history as a result of large-scale climate change events, such as when species recolonized much of North join communities, all resulting America after the last ice age. However, these changes were slower and in the generation of novel of a smaller magnitude than contemporary changes (Quintero and Wiens biotic communities, referred 2013), which are expected to continue under current global climate change to as community reassembly. projections (Huntley et al. 1997, McLachlan et al. 2005). Individual species Community reassembly can lead have different responses to climate change; some species will adjust via to changes in biodiversity, species phenotypic plasticity, some will adapt via evolutionary change, some species will leave communities (via range shifts or local extinctions), and immigrating interactions, trophic structure, species may join communities, all resulting in the generation of novel biotic and ecosystem processes and communities (community reassembly) (Møller et al. 2010). Community services. As well, community reassembly alters community composition and therefore can lead to changes reassembly brings novel groups in biodiversity, species interactions, trophic structure, and ecosystem of species into contact, introduces processes (Barry et al. 1995, Fritts and Rodda 1998, D’Antonio and Vitousek new predators, new diseases, and 1992, Nussey et al. 2005). As well, community reassembly brings novel groups of species into contact, introduces new predators, new diseases, new competitors into ecosystems, and new competitors into ecosystems, and can break down co-evolved and can break down co-evolved species interactions (Morgan et al. 2004, Brooker et al. 2007). For example, species interactions. Community the extinction of many vertebrates on the island of Guam is a result of their reassembly resulting from recent naïveté to a novel predator, the brown tree snake (Boiga irregularis), which climate change has already invaded the community (Fritts and Rodda 1998). been observed, including within Community reassembly resulting from recent climate change has already several bird communities in been observed within several bird communities (Lemoine et al. 2007, Europe and North America. Many Stralberg et al. 2009, Virkkala and Rajasärkkä 2011). In Europe, climate other changes are occurring or change has altered the composition of bird communities, with an increase expected. For example, in Ontario, in the proportion of long-distance migratory species and a decrease in the proportion of short-distance migratory species (Lemoine et al. 2007). Similarly, southern boreal forest tree species Stralberg et al. (2009) assessed the potential changes in the composition are expected to be gradually of California’s avian communities under future climate change scenarios. replaced by temperate forest They suggested that by 2070, species range shifts may lead to dramatic species as summer temperatures changes in the composition of California’s avian communities, such that warm, which will shift the dominant as much as 57% of the state may be occupied by novel communities. In herbivore species within the deer protected areas of Finland’s boreal forest, northern bird species have declined by 21% and southern species increased by 29%, coinciding with a rise in family (Cervidae) from moose to mean temperatures, and leading to a change in boreal community structure white-tailed deer, with the effects (Virkkala and Rajasärkkä 2011). Climate changes also appear to have altered potentially cascading to predator the bat communities of northern Costa Rica, as bat species are gradually species. colonizing higher elevations as the climate changes, and novel assemblages of bats now occur in the cloud forests (LaVal 2004). CLIMATE CHANGE RESEARCH REPORT CCRR-36 17

In Ontario, southern boreal forest tree species are expected to be gradually replaced by temperate forest species as summer temperatures warm, thereby changing the structure of present-day boreal forest communities (Galatowitsch et al. 2009). As the southern boreal forest is replaced by temperate plant species, it is expected that many temperate fauna will shift north as well. For example, the dominant herbivore species within the deer family (Cervidae) will shift from moose to white-tailed deer, which are expected to become abundant across Ontario (Frelich et al. 2012), with potentially cascading effects on predator species, such as grey wolves (Canis lupus) and eastern wolves (C. lycaon), as well as on the community’s food web as a whole, including other ungulate species such as caribou.

Community reassembly is expected to produce new and altered interactions among species (Tylianakis et al. 2007, 2008, Møller et al. 2010, Gilman et al. 2010). Species interactions can occur when their fundamental niches overlap (Schweiger et al. 2010), but not all interactions can be realized if the overlap of the fundamental niches of two species lies outside the current climate. However, as the climate changes, some of these interactions may become possible whereas others may disappear, changing the overall structure and functioning of communities (Schweiger et al. 2010).

Community reassembly may affect predator–prey interactions (a key process governing population dynamics; Murdoch et al. 2003) and modify fundamental food web properties (Møller et al. 2010). For example, community reassembly could be detrimental to predators if a specialist predator’s prey shifts its range outside of the predator’s community (Gilman et al. 2010). Conversely, reassembly might be beneficial to a species if it enables escape from antagonistic interactions, such as predation or competition. For example, species can benefit if they remain in their community while their predators and competitors shift their range to a new community (Menéndez et al. 2008, Van Grunsven et al. 2010). As well, if a novel prey expands its range into a new community, the prey base for predators in that community will increase (Gilman et al. 2010).

Individual plant and animal species will likely respond to climate change in different ways, shifting competitive balances to favour certain species over others (Tylianakis et al. 2008). Although novel species add to the species richness of a community upon their arrival, some can eventually cause the decline or even extinction of native species by out competing these species for limited resources, or via predation, disease, or replacement of resource species (D’Antonio and Dudley 1995, Dukes and Mooney 2004). Invading species often lack natural competitors or consumers and when released from their climatic constraints they can gain a competitive advantage in their expanded or introduced ranges thus significantly affecting communities (Dukes and Mooney 2004). For example, the red fire ant (Solenopsis invicta), an invasive species in the southern U.S., is extending its range north as the climate warms (Morrison et al. 2004). Invasive ants alter ecosystem processes by displacing native ant species that construct deep, long-lived nests rich in organic matter (MacMahon et al. 2000). As well, newly arriving competitors can take over available resources and prevent a later-arriving competitor from colonizing (Gilman et al. 2010).

With new species moving into communities, new diseases are expected to follow. Novel plants and animals can influence virus incidence in native species by introducing novel diseases and by increasing populations of vectors (D’Antonio and Meyerson 2002, Hampton et al. 2004, Malmstrom et al. 2005). The introduction of diseases to immunologically naïve hosts is often associated with increased prevalence and severity of disease (Bradley et al. 2005). Echinococcus multilocularis is a tapeworm that causes alveolar echinococcosis, a parasitic disease of canids and small rodents, which was previously unknown in northern Alaska (Bradley et al. 2005). The range expansion of the red fox to extreme northern Alaska may have had a role in the range expansion of E. multilocularis in brown lemmings (Lemmus trimucronatus) from the northern coast of Alaska (Bradley et al. 2005, Holt et al. 2005). Baylisascaris procyonis, a common roundworm of raccoons, is relatively harmless to raccoons, but can be fatal in rabbits, squirrels, groundhogs, other rodents, and humans (Kazacos 2001). Human infection by B. procyonis is an emerging health issue because raccoon populations are rapidly increasing, moving northward with climate change, and are living in close proximity to humans (Sorvillo et al. 2002, Bowman and Sadowski 2012). The parasite has been identified as one of the “deadly dozen” human pathogens thought to be affected by climate change (Wildlife Conservation Society 2008). 18 CLIMATE CHANGE RESEARCH REPORT CCRR-36

5.1 Breakdown of co-evolved interactions Although many species interactions have a long evolutionary history, this synchrony may be lost due to the relative speed of today’s anthropogenic climate change (Yurk and Powell 2009). Community reassembly is expected to disrupt co-evolved relationships between predators and their prey, plants and their pollinators, and others (Sherry et al. 2007; Tylianakis et al. 2007, 2008; Schweiger et al. 2010). In general, mutualistic interactions appear to be weakened by climate change (Tylianakis et al. 2008). For example, divergent range shifts in plants and pollinators are likely to change the amount of overlap in current species distribution, thereby disrupting mutualistic relationships (Schweiger et al. 2010). Changes in plant community composition and spatial mismatches in plant–pollinator responses to climate change may decrease pollinator availability for specialist plant species (Palmer et al. 2003). Similarly, Schweiger et al. (2008) modelled the climatic niche for the butterfly Boloria titania and its larval host plant Polygonum bistorta and found that the overlap of their climatic niches will be considerably reduced under future projected climate change scenarios, potentially disrupting this long-held trophic interaction. Walpole et al. (2012) demonstrated how unequal effects of increasing spring temperatures have led to an increase in the span of the breeding period for a community of anurans in Ontario. The asymmetric response by different anuran species may affect the type and strength of interspecific interactions (Donnelly and Crump 1998), and varying responses by species to climate change could alter the species composition of these communities and their fundamental ecological processes (Yang and Rudolf 2010).

5.2 Uncertainty The long-term ecological consequences of community reassembly and the resulting interactions among previously unknown combinations of species are difficult to determine. Predicting the effects of community reassembly is problematic because we often lack sufficient data to fully determine how species will respond to climate change or to predict how novel species may interact with one another. As well, the numerous abiotic and biotic factors that are potentially susceptible to climate change, the differential sensitivities to changing conditions among species, and the complexity of species interactions, make species- and community-specific projections difficult (Tylianakis et al. 2007).

Biological communities will not move as a unit; instead, differing influences on individual species will cause them each to move in their own direction and at their own rate (i.e., asymmetrically). We can, therefore, anticipate that current communities will disassemble and the individual species will assemble into novel communities; however, the specific composition of these novel communities cannot be accurately predicted. As well, the order in which novel species colonize a community is important in determining community composition (Connell and Slatyer 1977), and the timing of species colonization can lead to alternative compositions (Diamond 1975). Further, although climate is a major determinant of species distributions (Pearson and Dawson 2003, Luoto et al. 2007), other factors, such as habitat fragmentation (Opdam and Wascher 2004, Schweiger et al. 2010, Mantyka-Pringle et al. 2012) and invasive species (Walther et al. 2009, Mainka and Howard 2010), will interact with climate change to affect species distributions and the formation of novel communities in ways that are difficult to predict (i.e., synergies). For example, Rempel (2012) demonstrated how the effects of climate change on moose populations will be complex, involving main effects and interactions among numerous variables, such as summer heat stress, winter tick-induced death, brain worm, and predation. Another major unknown is how the strength of already established interactions will change. If predators shift their diets to novel prey, the distribution of strong and weak interactions within food webs will be rearranged. In addition, it is uncertain what new species interactions will occur and how strong these interactions will be (Lurgi et al. 2012).

The novel communities that result from climate change may persist as species adapt or coexist, or they may undergo even further change as species are excluded through competition, predation, or other biotic interactions (Stralberg et al. 2009). Some range shifts are expected to have cascading effects on community structure and the functioning of ecosystems (Lovejoy and Hannah 2005). Nevertheless, novel communities will be characterized by high levels of ecological change, and ecosystem functioning may differ in ways that we cannot yet predict (Stralberg et al. 2009). As such, these novel ecosystems will present challenges and opportunities for conservation and management; therefore, we should attempt to formally incorporate uncertainty into climate change research and assessment processes. CLIMATE CHANGE RESEARCH REPORT CCRR-36 19

5.3 Resilience Ecosystem resilience is the ability of an ecosystem to withstand and absorb disturbances, and to recover to its pre- disturbance state without losing function and services (Holling 1996, Willams et al. 2008, Cote and Darling 2010). The concept includes two separate processes: resistance (the degree of disturbance that causes a change in state), and recovery (the speed of return to the original state) (Tilman and Downing 1994, Holling 1996, Cote and Darling 2010). Resilience may be a fundamental factor contributing to the sustained production of natural resources and ecosystem services in communities faced with uncertainty (Gunderson and Holling 2002). The life history traits that are predicted to promote resilience and reduce extinction risk include high reproductive rates, fast life history, and short life span (McKinney 1997). Resilience is also affected by the size of ecosystems, as small, fragmented habitats reduce the likelihood that species will be able to maintain a viable population size in the face of shrinking optimal habitats (Williams et al. 2008). Meanwhile, the ability to disperse within and across habitats, the ability to track preferred climate envelopes, and the ability to rapidly expand following disturbance will depend on both reproductive rates and dispersal ability (Fjerdingstad et al. 2007). The resilience of ecosystems to changing environmental conditions is also determined by the biological diversity and genetic variability of species within the ecosystem (Rejmánek 1996, Peterson et al. 1998, Wilmers et al. 2002). Communities with lower species diversity or those lacking keystone species (Paine 1969, Power et al. 1996) may be more vulnerable to the effects of climate change than communities with higher diversity. The impacts of current climate change, especially interacting with other pressures such as habitat fragmentation, might be sufficient to overcome the resilience of even some large areas of primary forests, transforming them into a permanently changed state. The resulting ecosystem state may be poorer in terms of both biological diversity and delivery of ecosystem goods and services (Thompson et al. 2009).

5.4 Regime shifts The potential resilience of novel communities is generally unknown, although much research has been undertaken on this topic (Tilman and Downing 1994, Peterson et al. 1998). One common model argues that community resilience depends mostly on the number of species in the community (i.e., biodiversity; May 1973, Tilman 1999). Another model argues that resilience is an idiosyncratic product of the particular species present in the community (Lawton 1994).

In either case, the potential exists for ‘regime shifts’ to occur following community reassembly. Here, we define regime shifts after Folke et al. (2004), as alterations to ecosystem services that have consequent effects on human societies. A well-known example of a regime shift is a eutrophied lake, where high cyanobacteria counts and anoxic events lead to fish kills and a consequent loss of fishing opportunities (Folke et al. 2004). There is considerable potential for regime shifts in natural resources as a result of contemporary climate change (e.g., Chapin and Starfield 1997, Oosterkamp et al. 2000). As just one example, we are already seeing changes to furbearer distributions in the province that affect commercial fur harvesting activities (Koen et al. 2014). Widespread changes are also occurring in distributions of other animal and plant species (Varrin et al. 2007). It is likely that continued climate change will cause a variety of regime shifts in Ontario, altering socio-economically important ecosystem services. Regime shifts could occur as a gradual, continuous linear changes, or abruptly, as non-linear thresholds (Folke et al. 2004). 20 CLIMATE CHANGE RESEARCH REPORT CCRR-36

6. Recommendations

Resource managers in Ontario are faced with high uncertainty about Summary the future composition of natural communities, and about the potential Given high uncertainty about for deleterious regime shifts. Given these high levels of uncertainty, we recommend that decision-making processes be followed that allow for future biodiversity in Ontario, we learning. Folke et al. (2004) argued that in the face of high uncertainty, recommend implementing structured resilience can be built into natural systems through management that is decision-making processes, such flexible and open to learning. Adaptive management is an example of a as adaptive management, that allow structured decision-making process that explicitly accommodates learning for learning through management in the face of uncertainty. A key feature of the process is that management policies and actions are considered hypotheses that need to be evaluated activities to reduce future and compared to alternative hypotheses. Therefore, the adaptive uncertainties. We also recommend management process emphasizes creating and implementing different that such research and management policy options to facilitate learning through decision making, thereby actions be integrated at appropriate reducing uncertainty for future decisions. The learning process of adaptive management is often depicted as a loop (Figure 6). spatial and temporal scales. There are many opportunities to integrate research and management activities to reduce future uncertainties about the effects of climate change on terrestrial biodiversity. To provide just one example, climate warming and competition with coyotes have both been posited as processes leading to reduced lynx abundance at southern latitudes (Ripple et al. 2011; Koen et al. 2014). These alternatives could be evaluated by manipulating coyote harvest while controlling for differences in climate, and vice versa. Such an experiment could be done at little financial cost by collecting routine management data.

We also recommend that research and management be applied and integrated at appropriate spatial and temporal scales. Given the large spatial scale of climate change, we expect that many of the biodiversity changes will occur at large scales, such as at the ecoregional level, and this should be recognized in the application of management decisions.

Figure 6. A typical adaptive management loop. Adapted from Williams et al. (2009), and redrawn after MNR Risk Management (2013). CLIMATE CHANGE RESEARCH REPORT CCRR-36 21

We provide some specific suggestions for research and management activities below.

Research: • Conduct research to fill knowledge gaps about species, biotic interactions, and community responses to climate change. Integrate research findings with management decision making. • Due to the unpredictability of novel ecosystems, formally incorporate uncertainty into climate change research and assessment processes • Develop integrated monitoring programs linked to management to help detect and verify change as it occurs. This will help to guide strategic decision making and calibrate future modelling efforts. Such integrated monitoring should be done as part of MNR’s regular business. • Undertake long-term studies that can separate genetic from plastic components of adaptive responses. Long- term studies are also an important tool for understanding ecosystem change. • Research the mechanisms that confer community resilience to climate change (Williams et al. 2008). • Identify species, populations, and communities that require active human intervention to mitigate losses. • Develop models to better understand the complex potential outcomes of climate change on species, their interactions, and ecosystem functioning (Schmitz et al. 2003). • Evaluate potential synergies between climate change and other stressors such as invasive species, habitat fragmentations, and disease (McCarty 2001, Opdam and Wascher 2004). • Study genetic variability for fitness-related traits to identify species most at risk from climate change (Berteaux et al. 2004). • Further investigate the role of biodiversity in ecosystem structure and function. • Increase the monitoring of wildlife diseases and encourage collaboration between climate-change ecologists and infectious-disease researchers.

Management: • Given uncertainty about the exact nature of ecosystem responses to climate change, embrace strategic flexibility, characterized by risk-taking (including decisions of no action), capacity to reassess conditions frequently, and willingness to change course as conditions change (Hobbs et al. 2006). Flexibility will increase manager’s ability to deal with surprises as they occur (such as an insect pest suddenly switching from one generation per year to two generations per year, resulting in increased habitat damage). • Accept different levels of uncertainty and risk associated with planning at regional scales relative to local scales (Saxon et al. 2005). • Protect ecosystems with high biodiversity, especially those that maintain crucial components that may recover more easily from climatic disturbances, climate refugia, functional groups, keystone species, and multiple microhabitats within a biome. • Maintain connectivity across forest landscapes by reducing fragmentation, recovering lost habitats (forest types), expanding networks, and establishing buffer zones and ecological corridors (Thompson et al. 2009). • Restore ecosystem function and maintain or preserve natural ecosystem processes with minimal human interference. Ecosystem-based adaptation may require giving priority to some ecosys­tem services at the expense of others. • To promote ecosystem resilience, reduce and manage stresses faced by communities from other sources (such as habitat fragmentation, overharvest, invasive species, novel diseases) (Chapin et al. 2006). For example, minimize landscape fragmentation caused by road construction and urban development. • Move from a focus on species towards a focus on communities and landscapes as conservation and management approaches are updated to incorporate climate change (Groves et al. 2012). 22 CLIMATE CHANGE RESEARCH REPORT CCRR-36

7. Conclusions

Over the next 100 years, the average annual temperature in Ontario is expected to increase by 5 °C, with greater increases in winter than summer temperature (IPCC 2007, McKenney et al. 2010). Precipitation is expected to increase, and extreme weather events, such as drought, rain, hail and ice, and windstorms are expected to increase in frequency (IPCC 2007). In general, weather is expected to become more variable under climate change.

These changes will add to the other pressures already affecting Ontario’s biodiversity and ecosystem functioning. Although many species are thought to be able to cope with the direct effects of climate change, such as warming temperatures, indirect and interacting effects will likely play a larger role as climate change progresses (Callaghan et al. 2004, Luoto et al. 2007). Key drivers of these stresses are likely to be new synergistic interactions between climate change and other stressors, such as habitat loss, lack of connectivity, invasive species and disease, which are likely to constrain adaptive responses to climate change.

Globally, climate change is already significantly affecting species, biotic interactions, ecosystems, and the provision of ecosystem services. Changes in the timing of spring events (such as bud burst, flowering, migration, and breeding) have been widely documented (Parmesan and Yohe 2003, Root et al. 2003). Differing responses to climate change between interacting species has already resulted in increasing asynchrony in predator–prey and insect–plant systems,with mostly negative consequences, such as the decoupling of co-evolved species interactions between plants and their pollinators (Brooke et al. 2008, Post and Forchhammer 2008, Post et al. 2008). Species range shifts have also been well documented, as have expansions of warm-adapted communities (Chen et al. 2011, Hitch and Leberg 2007, Parmesan and Yohe 2003, Thomas et al. 2001). For example, species that were not historically adapted to Ontario’s climate, such as the Virginia opossum, have already begun to shift their ranges north into the province. As well, climate warming is contributing to the continuing range expansion of white-tailed deer, but is expected to lead to range contractions of moose and woodland caribou. Meanwhile, Ontario’s polar bear population, the southern-most population of polar bears in the world, may become extirpated within 45 years due to decreases in sea ice in Hudson Bay (Amstrup et al. 2007). Shifts in species distribution, phenology, abundance, and interactions can significantly alter community dynamics, leading to cascading effects throughout food webs and ecosystems (Coristine and Kerr 2011).

As well, invasive and non-native species, such as gypsy moth and mountain pine beetle, which were once restricted by colder winter temperatures, are expected to continue to spread at an increased rate (Mawdsley et al. 2009). Diseases and parasites (such as Lyme disease and raccoon roundworm) are also expected to spread, and shifts in abundances and ranges of parasites and their vectors are beginning to influence human disease dynamics (Pounds et al. 2006, van der Wal et al. 2008).

In southern Ontario and other areas with intensive land use and high levels of landscape fragmentation, the resulting barriers to population connectivity among habitat patches will likely affect species and communities through decreased dispersal (Wasserman et al. 2012), increased mortality (Fahrig et al. 1995), reduced genetic diversity (Reh and Seitz 1990, Wasserman et al. 2012), reduced recolonization following (Semlitsch and Bodie 1998), and ultimately may lead to population declines (Brown and Kodric-Brown 1977). Given that southern Ontario is one of the most species-rich areas of Canada, there is a clear need to escalate conservation efforts in these fragmented, human-dominated landscapes. CLIMATE CHANGE RESEARCH REPORT CCRR-36 23

Although evolutionary responses to climate change have been documented, there is little evidence that observed genetic shifts would be able to prevent predicted species losses. Abiotic changes affect each species in a community differently because each species has its own physiological optimum and experiences climate conditions differently (Gilman et al. 2010). As such, rapid, anthropogenic climate change is ultimately causing the re-shuffling of communities as species respond according to their unique individual niche requirements and dispersal capacities (Coristine and Kerr 2011). New communities and ecosystems will appear, and lead to changes in species interactions at both the species and the ecosystem level, as well as to changes in the provision of ecosystem services.

Projecting community and ecosystem responses to climate change is one of the major challenges in modern ecology (Warren et al. 2001, McRae et al. 2008, Mora et al. 2007). Responses to climate change vary considerably, depending on the species, species interactions, synergies between pressures, and the spatial and temporal scale considered (de Chazal and Rounsevell 2009). Therefore, it is impossible to accurately predict future circumstances of all variables, and our understanding of the ecological effects of global change remains limited because community- level changes have been poorly documented, in part, due to the paucity of long-term data and the complexity of numerous interacting effects.

Incorporating climate change effects into resource management requires an understanding of the risks posed by climate change, not only to individual species, but to ecological communities, ecosystems, and resource users as well. Rapid climate change could impose novel demands on species and community-level conservation efforts. As such, this report was developed to update stakeholders on recent research on community-level effects of climate change to help identify potential climate change vulnerabilities, and to aid in developing climate change action plans, strategies, and policies. As well, our hope is that this report will stimulate further research on community-level climate change effects, consideration of methods for adaptation and mitigation, and implementation of structured decision making to reduce future uncertainties. 24 CLIMATE CHANGE RESEARCH REPORT CCRR-36

References

Ahola, M., Laaksonen, T., Eeva, T. and E. Lehikoinen. 2007. Climate change can alter competitive relationships between resident and migratory birds. J. Anim. Ecol. 76: 1045–1052. Altizer, S., Harvell, D. and E. Friedle. 2003. Rapid evolutionary dynamics and disease threats to biodiversity. Trends Ecol. Evol. 18: 589–596. Amstrup, S.C., Marcot, B. G. and D.C. Douglas. 2007. Forecasting the rangewide status of polar bears at selected times in the 21st century. USDI, US Geol. Surv., Reston, VA. Anderson, J.T., Inouye, D.W., McKinney, A., Colautti, R. and T. Mitchell-Olds. 2012. Phenotypic plasticity and adaptive evolution contribute to advancing flowering phenology in response to climate change. Proc. R. Soc. London B. 279: 3843–3852. Austad, S.N. 1988. The adaptable opossum. Sci. Am. 258: 98–104. Ayala, F.J. 1982. Population and evolutionary genetics: a primer. The Benjamin/Cummings Publ. Co., Inc., Menlo Park, CA. Azevedo, J., Jack, S. Coulson, R. and D. Wunneburger. 2000. Functional heterogeneity of forest landscapes and the distribution and abundance of the red- cockaded woodpecker. For. Ecol. Manage. 127: 271–283. Barry, J.P., Baxter, C.H., Sagarin, R.D. and S.E. Gilman. 1995. Climate-related, long-term faunal changes in a California rocky intertidal community. Science 267: 672–675. Battisti, A., Stastny, M., Buffo, E. and S. Larsson. 2006. A rapid altitudinal range expansion in the processionary moth produced by the 2003 climatic anomaly. Glob. Change Biol. 12: 662–671. Beebee, T.J.C. 1995. Amphibian breeding and climate. Nature 374: 219–220. Bender, M.L., Sowers, T., Dickson, M.L., Orchado, J., Grootes, P., Mayewski,, P.A. and D.A. Meese. 1994. Climate connection between Greenland and Antarctica during the last 100,000 years. Nature 372: 663–666. Berg, M.P., Kiers, E.T., Driessen, G., Van der Heijden, M., Kooi, B.W., Kuenen, F., Liefting, M., Verhoef, H.A. and J. Ellers. 2010. Adapt or disperse: understanding species persistence in a changing world. Global Change Biol. 16: 587–598. Berger, S.A., Diehl, S., Stibor, H., Trommer, G., Ruhenstroth, M.,Jager, C. and M. Striebel. 2007. Water temperature and mixing depth affect timing and intensity of events during spring succession of the plankton. Oecologia 150: 643–654. Berteaux, D., Réale, D., McAdam, A.G. and S. Boutin. 2004. Keeping pace with fast climate change: can Arctic life count on evolution? Integr. Comp. Biol. 44: 140–151. Berthold, P., Fiedler, W., Schlenker, R. and U. Querner. 1998. 25-year study of the population development of Central European songbirds: a general decline, most evident in long-distance migrants. Naturwissenschaften 85: 350–353. Berthold, P., Helbig, A.J., Mohr, G. and U. Querner. 1992. Rapid microevolution of migratory behaviour in a wild bird species. Nature 360: 668–670. Blaustein, A.R., Belden, L.K., Olson, D.H., Green, D.M., Root, T.L. and J.M. Kiesecker. 2001. Amphibian breeding and climate change. Conserv. Biol. 15: 1804–1809. Both, C. 2007. Comment on ‘Rapid advance of spring arrival dates in long-distance migratory birds’. Science 315: 598 Both, C. and M.E. Visser. 2001. Adjustment to climate change is constrained by arrival date in a long-distance migrant bird. Nature 411: 296–298. Both, C., Bouwhuis, S., Lessells, C.M. and M.E. Visser. 2006. Climate change and population declines in a long-distance migratory bird. Nature 441: 81–83. Botkin, D.B., Woodby, A.D. and A.R. Nisbet. 1991. Kirtland’s warbler habitats: A possible early indicator of climatic warming. Biol. Conserv. 56: 63–78. Bowman, J. and C. Sadowski. 2012. Vulnerability of furbearers in Ontario’s Clay Belt to climate change. Ont. Min. Nat. Resour., Appl. Res. Devel. Br., Sault Ste. Marie, ON. Clim. Change Res. Rep. CCRR-25. 11 p. Bowman, J., Holloway, G.L., Malcolm, J.R., Middel, K.R. and P.J. Wilson. 2005. Northern range boundary dynamics of southern flying squirrels: evidence of an energetic bottleneck. Can. J. Zool. 83: 1486–1494. Bradley, C. and S. Altizer. 2005. Parasites hinder monarch butterfly flight ability: implications for disease spread in migratory hosts. Ecol. Lett. 8: 290–300. Bradley, M.J., Kutz, S.J., Jenkins E. and T.M. O’Hara. 2005. The potential impact of climate change on infectious diseases of Arctic fauna. Int. J. Circumpolar Health 64: 468–477. Bradley, N.L., Leopold, A.C., Ross, J. and W. Huffaker. 1999. Phenological changes reflect climate change in Wisconsin. Proc. Nat. Acad. Sci. USA 96: 9701–9704. Bradshaw, W.E. and C.M. Holzapfel. 2007. Tantalizing timeless. Science 316: 1851–1852. Brooke, M. de L., Butchart, S.H.M., Garnett, S.T., Crowley, G.M., Mantilla-Beniers, N.B. and A. J. Stattersfield. 2008. How fast are threatened bird species moving towards extinction? Conserv. Biol. 22: 417–427. Brooker, R.W., Travis, J.M.J, Clarke, E.J. and C. Dytham. 2007. Modelling species’ range shifts in a changing climate: the impacts of biotic interactions, dispersal distance and rate of climate change. J.Theor. Biol. 245: 59–65. CLIMATE CHANGE RESEARCH REPORT CCRR-36 25

Brown J.H. and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58: 445–49. Butler, C.J. 2003. The disproportionate effect of global warming on the arrival dates of short-distance migratory birds in North America. Ibis 145: 484–495. Callaghan, T.V., et al. 2004. Response to projected changes in climate and UV-B at the species level. Ambio 33:418–435. Candau, J.N. and R.A. Fleming. 2008. Forecasting the response to climate change of the major natural biotic disturbance regime in Ontario’s forests: the spruce budworm. Ont. Min. Nat. Resour., Appl. Res. Devel. Br., Sault Ste. Marie, ON. Clim. Change Res. Rep.CCRR-13. 14 p. Cardillo, M., Mace, G.M., Gittleman, J.L. and A. Purvis. 2006. Latent extinction risk and the future battlegrounds of mammal conservation. Proc. Nat. Acad. Sci. USA 103: 4157–4161. Carlini, A.R., Coria, N.R., Santos, M.M. and S.M. Buján. 2005. The effect of chinstrap penguins on the breeding performance of Adélie penguins. Folia Zool. 54: 147−158. Carr, D., Bowman, J., Kyle, C.J., Tully, S.M., Koen, E.L., Robitaille, J.-F. and P. J. Wilson. 2007a. Rapid homogenization of multiple sources: genetic structure of a recolonizing population of fishers. J. Wildl. Manage. 71: 1853−1861. Carr, D., Bowman, J. and P. J. Wilson. 2007b. Density-dependent dispersal suggests a genetic measure of habitat suitability. Oikos 116: 629−635. Carroll, C. 2007. Interacting effects of climate change, landscape conversion, and harvest on carnivore populations at the range margin: Marten and Lynx in the northern Appalachians. Conserv. Biol. 21: 1092−1104. Chapin III, F.S. and Starfield, A.M. 1997. Time lags and novel ecosystems in response to transient climatic change in Arctic Alaska. Climatic Change 35: 449–461. Chapin III, F.S., et al. 2006. Policy strategies to address sustainability of Alaskan boreal forests in response to a directionally changing climate. Proc. Nat. Acad. Sci. USA 103: 16637–16643. Chen, I.-C., Hill, J.K., Ohlemüller, R., Roy, D.B. and C.D. Thomas. 2011. Rapid range shifts of species associated with high levels of climate warming. Science 333: 1024−1026. Cobben, M.M.P., Verboom, J., Opdam, P.F., Hoekstra, R.F., Jochem, R. and M.J. Smulders. 2012. Wrong place, wrong time: climate change-induced range shift across fragmented habitat causes maladaptation and declined population size in a modelled bird species. Global Change Biol. 18: 2419−2428. Colombo, S.J., McKenney, D.W., Lawrence, K.M. and P.A. Gray. 2007. Climate change projections for Ontario: Practical information for policymakers and planners. Ont. Min. Nat. Resour., Appl. Res. Devel. Br., Sault Ste. Marie, ON. Clim. Change Res. Rep. CCRR-05. 38 p. Connell, J.H. and R.O. Slatyer. 1977. Mechanisms of succession in natural communities and their role in community stability and organization. Am. Nat. 111: 1119−l 144. Corbet, S.A. 2000. Conserving compartments in pollination webs. Conserv. Biol. 14: 1229–1231. Coristine, L.E. and J.T. Kerr. 2011. Habitat loss, climate change, and emerging conservation challenges in Canada. Can. J. Zool. 89:435−451. Corser, J.D. 2001. Decline of disjunct green salamander (Aneides aeneus) populations in the Southern Appalachians. Biol. Conserv. 97: 119–126. Cote, I. and E. Darling. 2010. Rethinking ecosystem resilience in the face of climate change. PLOS Biol 8(7): e1000438. Curry, R.L. 2005. Hybridization in chickadees: much to learn from familiar birds. Auk 122: 747–758. Cushing, D.H. 1990. Plankton production and year-class strength in fish populations: an update of the match/mismatch hypothesis. Adv. Marine Biol. 26: 249–293. D’Antonio, C.M. and L.A. Meyerson. 2002. Exotic plant species as problems and solutions in ecological restoration: a synthesis. Restor. Ecol. 10: 703–713. D’Antonio, C.M. and P.M. Vitousek. 1992. Biological invasions by exotic grasses, the grass/fire cycle, and global change. Ann. Rev. Ecol. Syst. 23: 63–87. D’Antonio, C.M. and T.L. Dudley. 1995. Biological invasions as agents of change on islands versus mainlands. In: Vitousek, P.M., Loope, L.L.and H. Adsersen (eds) Islands: Biological Diversity and Ecosystem Function. Springer, Berlin Heidelberg New York, pp 103–121. Davis, M.B. 1986. Climatic instability, time lags and community disequilibrium. Pp. 269–284 in Diamond, J. and T. Case (eds.). Community Ecology. Harper & Row Publishers Inc., Cambridge, UK. de Chazal, J. and M.D.A. Rounsevell. 2009. Land-use and climate change within assessments of biodiversity change: A review. Global Environ. Change 19: 306–315. de Jong, G. 2005. Evolution of phenotypic plasticity: pattern of plasticity and the emergence of ecotypes. New Phytol.166: 101–108. de la Rocque, S., Rioux, J.A. and J. Slingenbergh. 2008. Climate change: Effects on animal disease systems and implications for surveillance and control. Rev. Sci. Tech. 27: 339–354. Debinski, D.M. and R.D. Holt. 2000. A survey and overview of habitat fragmentation experiments. Conserv. Biol. 14: 342–355. Derocher, A.E., Lunn, N.J. and I. Stirling. 2004. Polar bears in a warming climate. Integr. Comp. Biol. 44: 163–176. DeWitt, T.J., Sih, A. and D.S. Wilson. 1998. Costs and limits of phenotypic plasticity. Trends Ecol. Evol. 13: 77–81. Diamond, J.M. 1975. Assembly of species communities. Pp 342–444in Cody, M.L.and J.M. Diamond (eds). Ecology and Evolution of Communities. Harvard University Press, Cambridge, MA. 26 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Dickey, M. H., Gauthier, G. and M.C. Cadieux. 2008. Climatic effects on the breeding phenology and reproductive success of an arctic-nesting goose species. Global Change Biol. 14: 1973–1975. Donnelly, M.A. and M.L. Crump. 1998. Potential effects of climate change on two Neotropical amphibian assemblages. Climate Change. 39: 541–561. Dukes, J.S. and H.A. Mooney. 2004. Disruption of ecosystem processes in western North America by invasive species. Rev. Chil. Hist. Nat. 77: 411–437. Dunn, P.O. and D.W. Winkler. 1999. Climate change has affected the breeding date of tree swallows throughout North America. Proc. R. Soc. Lond. B 266: 2487–2490. Edwards, M. and A.J. Richardson. 2004. The impact of climate change on the phenology of the plankton community and trophic mismatch. Nature 430: 881–884. Elzinga, J.A., Atlan, A., Arjen, B., Gigord, L., Weis, A.E. and G. Bernasconi. 2007. Time after time: flowering phenology and biotic interactions. Trends in Ecology and Evolution 22: 432–439. Environment Canada. 2013. Climate trends and variations bulletin – Annual 2012. https://www.ec.gc.ca/adsc-cmda/default.asp?lang=En&n=4A21B114-1 [accessed 31 October 2013]. Epps, C.W., Palsboll, P.J., Wehausen, J.D., Roderick, G.K. and D.R. McCullough. 2006. Elevation and connectivity define genetic refugia for mountain sheep as climate warms. Mol. Ecol. 15: 4295–4302. Epstein, P.R. 2001. Climate change and emerging infectious diseases. Microbes Infect 3: 747–754. Fahrig, L. 1997. Relative effects of habitat loss and fragmentation on species extinction. J. Wildl. Manage. 61: 603–610. Fahrig, L. 2001. How much habitat is enough? Biol. Conserv. 100: 65–74. Fahrig, L. 2003. Effects of habitat fragmentation on biodiversity. Ann. Rev. Ecol. Syst. 34: 487–515. Fahrig, L., Pedlar, J.H., Pope, S.E., Taylor, P.D. and J.F. Wegner. 1995. Effect of road traffic on amphibian density. Biol. Conserv. 73:177–182. Fenton, A. and M. Spencer. 2010. Linking population, community and ecosystem ecology within mainstream ecology. Pp. 19–39 in Raffaelli, D. and C.L.J. Frid (eds.). Ecosystem Ecology, A New Synthesis .Cambridge University Press. Cambridge, UK: Fitter, A.H. and R.S.R. Fitter. 2002. Rapid changes in flowering time in British plants. Science 296: 1689–1691. Fjerdingstad, E.J., Schtickzelle, N., Manhes, P., Gutierrez, A. and J. Clobert. 2007. Evolution of dispersal and life history strategies: Tetrahymena ciliates. BMC Evol. Biol. 7: 133. Fleming, R.A. 2000. Climate change and forest disturbance regimes in Canada’s boreal forests. World Resour. Rev. 12: 520–554. Folke, C., Carpenter, S.R., Walker, B.H., Scheffer, M., Elmqvist, T., Gunderson, L.H., and C.S. Holling. 2004. Regime shifts, resilience and biodiversity in ecosystem management. Ann. Rev. Ecol. Evol. Syst. 35: 557–581. Forchhammer, M.C., Post, E. and N.C. Stenseth. 1998. Breeding phenology and climate. Nature 391: 29–30. Forchhammer, M.C., Post, E., Berg, T.B.G, Høye, T.T. and N.-M. Schmidt. 2005. Large-scale climatic fingerprint in local short-term plant and herbivore behaviour. Ecology 86: 2644–2651. Forchhammer, M.C., Post, E., Stenseth, N.C. and D.M. Boertmann. 2002. Long-term responses in arctic ungulate dynamics to changes in climatic and trophic processes. Pop. Ecol. 44:113–120. Frankham, R. 1999. Quantitative genetics in . Gen. Res.74: 237–244. Frazer, N.B., Greene, J.L. and J.W. Gibbons. 1993. Temporal variation in growth rate and age at maturity of male painted turtles, Chrysemys picta. Am. Midl. Nat. 130: 314–324. Frelich, L.E., Peterson, R.O., Dovčiak, M., Reich, P.B., Vucetich, J.A. and N. Eisenhauer. 2012. Trophic cascades, invasive species, and body size hierarchies interactively modulate climate change responses of ecotonal temperate boreal forest. Phil. Trans. R. Soc. Lond. B 367: 2955–2961. Frey, J.K. 1992. Response of a mammalian faunal element to climatic changes. J. Mammal. 73: 43–50. Fritts, T.H. and G.H. Rodda. 1998. The role of introduced species in the degradation of island ecosystems. Ann. Rev. Ecol. Syst. 29: 113–140. Futuyma, D. 1998. Evolutionary Biology (3rd ed). Sinaeur Associates, Sunderland, MA. Galatowitsch, S., Frelich, L. and L. Phillips-Mao. 2009. Regional climate change adaptation strategies for biodiversity conservation in a midcontinental region of North America. Biol. Conserv. 142: 2012–2022. Galley, K. (ed.). 2004. Global climate change and wildlife in North America. The Wildlife Society Technical Review. 04-2. Garbrecht, J.D. and T.C. Piechota. 2006. Water resources and climate. Pp. 19–33 in J.D. Garbrecht and T.C. Piechota (eds.). Climate Variations, Climate Change, and Water Resources Engineering; American Society of Civil Engineers, Reston, VA.. Garroway C.J. and H.G. Broders. 2005.The quantitative effects of population density and winter weather on the body condition of white-tailed deer (Odocoileus virginianus) in Nova Scotia, Canada. Can. J. Zool. 83: 1246–1256. Garroway, C.J., Bowman, J., Cascaden, T.J., Holloway, G.L., Mahan, C.J., Malcolm, J.R., Steele, M.A., Turner, G. and P.J. Wilson. 2010. Climate change induced hybridization in flying squirrels. Global Change Biol. 16: 113–121. CLIMATE CHANGE RESEARCH REPORT CCRR-36 27

Gatewood, A.G., et al. 2009. Climate and tick seasonality are predictors of Borrelia burgdorferi genotype distribution. Appl. Environ. Microbiol. 75: 2476–2483. Gibbs, J.P. and A.R. Breisch. 2001. Climate warming and calling phenology of frogs near Ithaca, New York, 1900-1999. Conserv. Biol. 15: 1175–1178. Gibbs, J.P. and N.E. Karraker. 2005. Effects of warming conditions in eastern North American forests on red-backed salamander morphology. Conserv. Biol. 20: 913–917. Gilman, S.E., Urban, M.C., Tewksbury, J., Gilchrist G.W. and R.D. Holt. 2010. A framework for community interactions under climate change. Trends Ecol. Evol. 25: 325–331. Gilpin, M.E. and M.E. Soule. 1986. Minimum viable populations: processes of species extinction. Pp. 19–34 in Soule, M.E. (ed.). Conservation Biology: The Science of Scarcity and Diversity. Sinauer, Sunderland, MA. Grant, P.R. and B.R. Grant. 2010. Conspecific versus heterospecific gene exchange between populations of Darwin’s finches. Phil. Trans. R. Soc. Lond. B365: 1065–1076. Gray, P.A. 2005. The impacts of climate change on diversity in forested ecosystems: some examples. For. Chron. 81: 655–661. Greenbank, D.O. 1963. Climate and the spruce budworm. Memoirs Entomol. Soc. Can. 31: 174– 180. Greer, A., Ng, V. and D. Fisman. 2008. Climate change and infectious diseases in North America: the road ahead. Can. Med. Assoc. J. 178: 6. Gritti, E.S., Smith, B. and M.T. Sykes. 2006. Vulnerability of mediterranean basin ecosystems to climate change and invasion by exotic plant species. J. Biogeogr. 33: 145–157. Groves, C.R., Game, E.T., Anderson, M.G., Cross, M., Enquist, C., Ferdaña, Z., Girvetz, E., Gondor, A., Hall, K.R. and J. Higgins. 2012. Incorporating climate change into systematic conservation planning. Biodivers. Conserv. 21: 1651–1671. Gunderson, L.H. and C.S. Holling. 2002. Panarchy: Understanding Transformations in Human and Natural Systems. Island Press, Washington, DC. Gunn, A. and T. Skogland. 1997. Responses of caribou and reindeer to global warming. Ecol. Studies 124: 189–200. Gylfe, A.,Bergstrom, S.,Lundstrom, J. and B. Olsen. 2000. Reactivation of Borrelia infection in birds.Nature 403:724–725. Hampton, J., Kleiber, P., Langley, A. and K. Hiramatsu. 2004. Stock assessment of bigeye tuna in the western and central Pacific Ocean. SCTB17. WP- SA-2. Hanski, I. and M. Gilpin. 1991. Metapopulation dynamics: a brief history and conceptual domain. Biol. J. Linn. Soc. 42: 3–16. Hardin, G. 1960. The competitive exclusion principle. Science 131: 1292–1297. Harrington, R., et al. 2007. Environmental change and the phenology of European aphids. Glob. Change Biol. 13: 1550–1564. Harrington, R., Woiwod, I. and T. Sparks. 1999. Climate change and trophic interactions. Trends Ecol. Evol. 14: 146–150. Harvell, C.D., Altizer, S., Cattadori, I.M., Harrington, L. and E. Weil. 2009. Climate change and wildlife disease: when does the host matter most? Ecology 90: 912–920. Harvell, C.D., Mitchell, C.E., Ward, J.R., Altizier, S., Dobson, A.P., Ostfeld, R.S. and M.D. Samuel. 2002. Climate warming and disease risks for terrestrial and marine biota. Science 296: 2158–2162. Haydon, D.T., Stenseth, N.C., Boyce, M.S. and P.E. Greenwood. 2001. Phase coupling and synchrony in the spatiotemporal dynamics of muskrat and mink populations across Canada. Proc. Nat. Acad. Sci. USA. 98: 13149–13154. Hellmann, J.J., et al. 2008. Five potential consequences of climate change for invasive species. Conserv. Biol. 22: 534–543. Hersteinsson, P. and D.W. Macdonald. 1992. Interspecific competition and the geographical distribution of red and Arctic foxes, Vulpes vulpes and Alopex lagopus. Oikos 64: 505–515. Hitch, A.T. and P.L. Leberg. 2007. Breeding distributions of North American bird species moving north as a result of climate change. Conserv. Biol. 21: 534–539. Hobbs, R.J., et al. 2006. Novel ecosystems: theoretical and management aspects of the new ecological world order. Global Ecol.Biogeogr. 15: 1–7. Hoberg, E.P., Polley, L., Jenkins, E.J., Kutz, S.J., Veitch, A.M. and B.T. Elkin. 2008. Integrated approaches and empirical models for investigation of parasitic diseases in northern wildlife. Emerg. Infect. Dis. 14: 10–17. Hoffmann, A.A. and C.M. Sgro. 2011. Climate change and evolutionary adaptation. Nature 470: 479–485. Holling, C.S. 1996. Engineering resilience versus ecological resilience. Pp. 31–44 in Schulze, P. (ed.). Engineering Within Ecological Constraints. National Academy Press, Washington, DC. Holt, D.W., Hanns, C., O’Hara, T., Burek, K. and R. Frantz. 2005. New distribution records of Echinococcus multilocularis in the brown lemming from Barrow, Alaska, USA. J. Wildl. Dis. 41: 257–259. Holt, R.D. 1990. The microevolutionary consequences of climate change. Trends Ecol. Evol. 5: 311–315. Houghton, J.T., et al. 2001. Climate change 2001: the scientific basis. Contribution of working group 1 to the third assessment report of the intergovernmental panel on climate change. Cambridge University Press, Cambridge, UK. 28 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Humphries, M.M., Thomas, D.W. and J.R. Speakman. 2002. Climate-induced energetic constraints on the distribution of hibernating mammals. Nature 418: 313–316. Humphries, M.M., Umbanhowar, J. and K.S. McCann. 2004. Bioenergetic prediction of climate change impacts on northern mammals. Integr. Comp. Biol. 44: 152–162. Hunter, C.M., Caswell, H., Runge, M.C., Regeher, E.V., Amstrup, A.C. and I. Stirling. 2010. Climate change threatens polar bear populations: a stochastic demographic analysis. Ecology 91: 2883–2897. Huntley, B., Cramer, W., Morgan, A.V., Prentice, H.C. and J.R.M. Allen. 1997. Past and Future Rapid Environmental Changes: The Spatial and Evolutionary Responses of Terrestrial Biota. NATO ASI Series, Series 1, Springer-Verlag, Berlin, Germany. Hussell, D.J.T. 2003. Climate change, spring temperatures, and timing of breeding of tree swallows (Tachycineta bicolour) in southern Ontario. Auk 120: 607–618. Inouye, D.W., B. Barr, Armitage, K. B. and B.D. Inouye. 2000. Climate change is affecting altitudinal migrants and hibernating species. Proc. Nat. Acad. Sci. USA 97: 1630–1633. [IPCC] Intergovernmental Panel on Climate Change. 2006. 2006 IPCC guidelines for national greenhouse gas inventories. Prepared by the National Greenhouse Gas Inventories Programme. IPCC-IGES, Japan. [IPCC] Intergovernmental Panel on Climate Change. 2007. Climate change 2007: the physical science basis: summary for policymakers. A Contribution of Working Group I to the Fourth Assessment Report of the IPCC, Geneva, Switzerland. Janzen, F.J. 1994. Climate change and temperature-dependent sex determination in reptiles. Proc. Nat. Acad. Sci. USA 91: 7487–7490. Jetz, W., Wilcove, D. and A. Dobson. 2007. Projected impacts of climate and land-use change on the global diversity of birds. PLOS Biology 5: e157. Jonzen, N., et al. 2006. Rapid advance of spring arrival dates in long-distance migratory birds. Science 312: 1959–1961. Kaeslin, E., Redmond, I. and N. Dudley. (eds.). 2012. Wildlife in a changing climate. Food and Agriculture Organization of the United Nations.FAO Forestry Paper 167. Rome, IT. Kanda, L.L. 2005. Winter energetics of Virginia opossums Didelphis virginiana and implications for the species’ northern distributional limit. Ecography 28: 731–744. Kareiva, P.M., Kingsolver, J.G. and R.B. Huey. 1993. Biotic Interactions and Global Change. Sinauer Associates Inc., Sunderland, MA. Kazacos, K.R. 2001. Baylisascaris procyonis and related species. Pp. 301-341 in Samuel, W.M., M.J. Pybus and A.A. Kocan (eds.). Parasitic Diseases of Wild Mammals, 2nd ed. Iowa State University Press, Ames, IA. Kelly, B.P., et al. 2010. Status review of the ringed seal (Phoca hispida). NOAA Tech. Mem. NMFS-AFSC-212. Kerr, J. and L. Packer. 1998. The impact of climate change on mammal diversity in Canada. Environ. Mon. Assess. 49: 263270. Kerr, J.T. and J. Cihlar. 2004. Patterns and causes of species endangerment in Canada. Ecol. Appl. 14: 743–753. Koen, E.L., Bowman, J., Murray, D.L. and P.J. Wilson. 2014. Climate change reduces genetic diversity of Canada lynx at the trailing range edge. Ecography 10.1111/j.1600-0587.2013.00629.x. Kotiaho, J.S., Kaitala, V., Komonen, A. and J. Paivinen. 2005. Predicting the risk of extinction from shared ecological characteristics. Proc. Nat. Acad. Sci. USA102: 1963–1967. Krohn, W.B., Elowe, K.D. and R.B. 1995. Relations among fishers, snow, and martens: development and evaluation of two hypotheses. For. Chron. 71: 97–106. Krohn, W.B., Zielinski, W.J. and R.B. Boone. 1997. Relations among fishers, snow, and martens in California: results from small-scale spatial comparisons. Pp. 211-232 in Proulx, G., Bryant, H.N. and Woodard, P.M. (eds.). Martes: Taxonomy, Ecology, Techniques, and Management. Prov. Mus. Alberta, Edmonton, AB. Kruuk, L.E.B., Slate, J., Pemberton, J.M, Brotherstone, S., Guinness, F., and T. Clutton-Brock. 2002. Antler size in red deer: heritability and selection but no evolution. Evolution 56: 1683–1695. Kuhn, K., Campbell-Lendrum, D., Haines, A. and J. Cox. 2005. Using Climate to Predict Infectious Disease Epidemics. World Health Organization, Geneva. Kutz, S.J., Hoberg, E.P., Polley, L. and E.J. Jenkins. 2005. Global warming is changing the dynamics of Arctic host-parasite systems. Proc. R. Soc. Lond. B 272: 2571-2576. Kutz, S.J., Veitch, A.M., Hoberg, E.P., Elkin, B.T., Jenkins, E.J. and L. Polley. 2001. New host and geographic records for two protostrongylids in Dall’s sheep. J. Wildl. Dis. 37: 761–774. Lande, R. and S. Shannon. 1996. The role of genetic variation in adaptation and population persistence in a changing environment. Evolution 50:434–437. Larivière, S. 2004. Range expansion of raccoons in the Canadian prairies: review of hypotheses. Wildl. Soc. Bull. 32: 955–963. LaVal, R.K. 2004. Impact of global warming and locally changing climate on tropical cloud forest bats. J. Mammal. 85: 237–244. Lawton, J.H. 1994. What do species do in ecosystems? Oikos 71: 367–374. Le Conte, Y. and M. Navajas. 2008. Climate change: impact on honey bee populations and diseases. Rev. Sci. Tech. Off. Internat. Epizooties 27: 499–510. CLIMATE CHANGE RESEARCH REPORT CCRR-36 29

Lemoine, N. and K. Böhning-Gaese. 2003. Potential impact of global climate change on species richness of long-distance migrants. Conserv. Biol. 17: 577–586. Lemoine, N., Bauer, H.-G., Peintinger, M. and K. Böhning-Gaese. 2007. Impact of land-use and global climate change on species abundance in a Central European bird community. Conserv. Biol.21: 495–503. Lenarz, M.S. Nelson, M.E., Schrage, M. and A.J. Edwards. 2009. Temperature mediated moose survival in northeastern Minnesota. J. Wildl. Manage. 73: 503–510. Lenoir, J., Gegout, J.C., Marquet, P.A., de Ruffray, P. and H. Brisse. 2008. A significant upward shift in plant species optimum elevation during the 20th century. Science 320: 1768–1771. Lindgren, E. and R. Gustafson. 2001. Tick-borne encephalitis in Sweden and climate change. Lancet 358: 16–18. Lishman, G.S. 1985. The comparative breeding biology of Adélie and chinstrap penguins Pygoscelis adeliae and P. antarctica at Signy Island, South Orkney Islands. Ibis 127: 84−99 Littell, J.S., Elsner, M.M., Mauger, G., Lutz, E., Hamlet, A.F. and E. Salathe. 2011. Regional climate and hydrologic change in the northern US Rockies and Pacific Northwest: internally consistent projections of future climate for resource management. University of Washington, Seattle, WA. Litvaitis, J.A. 1992. Niche relations between coyotes and sympatric Carnivora. Pages 73–85 in A. H. Boer, (ed.). Ecology and Management of the Eastern Coyote. University of New Brunswick Wildlife Research Unit, Fredericton, NB. Loehle, C. 1995. Social barriers to pathogen transmission in wild animal populations. Ecology 76: 326–335. Logan, J.A. and J.A. Powell. 2001. Ghost forests, global warming, and the mountain pine beetle. Am. Entomol. 47: 160–173. Lovejoy, T.E. and L. Hannah. 2005. Climate Change and Biodiversity. Yale University Press. New Haven, CT. Luoto. M., Virkkala, R. and R.K. Heikkinen. 2007. The role of land cover in bioclimatic models depends on spatial resolution. Global Ecol. Biogeogr. 16: 34–42. Lurgi, M., López, B.C. and J.M. Montoya. 2012. Novel communities from climate change. Phil.Trans. R. Soc. B367: 2913–2922. Lusseau, D., Williams, R. Wilson, B., Grellier, K., Barton, T.R., Hammond P.S. and P.M. Thompson. 2004. Parallel influence of climate on the behaviour of Pacific killer whales and Atlantic bottlenose dolphins. Ecol. Lett. 7: 1068–1076. Lynch, H.J., Fagan, W.F., Naveen, R., Trivelpiece, S.G. and W.Z. Trivelpiece. 2012. Differential advancement of breeding phenology in response to climate may alter staggered breeding among sympatric pygoscelid penguins. Mar. Ecol. Prog. Ser. 454: 135−145 Lynch, H.J., Naveen, R. and W.F. Fagan. 2008. Censuses of penguins, blue-eyed shags, and southern giant petrel populations in the Antarctic peninsula, 2001–2007. Mar. Ornithol. 36: 83−97. [MEA] Millenium Assessment. 2005. Millennium Ecosystem Assessment. Ecosystems and Human Well-being: Current State and Trends. Vol. 1, Washington, Covelo, London.http://www.unep.org/maweb/en/Framework.aspx#download MacArthur, R.H. and R. Levins. 1967. The limiting similarity, convergence and divergence of coexisting species. Am. Nat.101: 377–385. MacDougall, A.S. and R. Turkington. 2005. Are invasive species the drivers or passengers of change in degraded ecosystems? Ecology 86:42–55. MacInnes, C.D., Dunn, E.H., Rusch, D.H., Cooke, F. and F.G. Cooch. 1990. Advancement of goose nesting dates in the Hudson Bay region 1951-1986. Can. Field-Nat. 104: 295–297. MacMahon, J.A., Mull, J.F. and T.O. Crist. 2000. Harvester ants (Pogonomyrmex spp.): their community and ecosystem influences. Ann. Rev. Ecol. Syst. 31:265–295. Mainka, S.A. and G.W. Howard. 2010. Climate change and invasive species: double jeopardy. Integr. Zool. 5:102-111. Malcolm, J.R., Markham, A. and R.P. Neilson. 2002. Estimated migration rates under scenarios of global climate change. J. Biogeogr. 29: 835–849. Mallory, M.L., Gilchrist, H.G., Fontaine, A.J. and J.A. Akearok. 2003. Local ecological knowledge of ivory gull declines in Arctic Canada. Arctic 56: 293–298. Malmstrom, R.R., Cottrell, M.T., Elifantz, H. and D.L. Kirchman. 2005. Biomass production and assimilation of dissolved organic matter by SAR11 bacteria in the northwest Atlantic Ocean. Appl. Environ. Microbiol. 71: 2979–2986. Mantyka-Pringle, C., Martin, T.G. and J.R. Rhodes. 2012. Interactions between climate and habitat loss effects on biodiversity: a systematic review and meta-analysis. Global Change Biol. 18: 1239–1252. Mattson, W.J. and R.A. Haack. 1987. The role of drought in outbreaks of plant-eating insects. Bioscience 37: 110–118. Mawdsley, J.R., O’Malley, R. and D.S. Ojima. 2009. A review of climate-change adaptation strategies for wildlife management and biodiversity conservation. Conserv. Biol. 23: 1080–1089. May, R. M. 1973. Qualitative stability in model ecosystems. Ecology 54: 638–641. McCarty, J.P. 2001. Ecological consequences of recent climate change. Conserv. Biol. 15: 320–331. McKenney, D.W., Pedlar, J.H., Lawrence, K., Campbell, K. and M.F. Hutchinson. 2007. Potential impacts of climate change on the distribution of North American tree species. Bioscience 57: 929–937. 30 CLIMATE CHANGE RESEARCH REPORT CCRR-36

McKenney, D.W., Pedlar, J.H., Lawrence, K., Gray, P.A., Colombo, S.J. and W.J. Crins. 2010. Current and projected future climatic conditions for ecoregions and selected natural heritage areas in Ontario. Ont. Min. Nat. Resour., Appl. Res. Devel. Br., Sault Ste. Marie, ON. Clim. Change Res. Rep. CCRR-16. 43 p. + CD. McKinney, M.L. 1997. Extinction vulnerability and selectivity: combining ecological and paleontological views. Ann. Rev. Ecol. Syst. 28: 495–516 McLachlan, J.S., Clark, J.S. and P.S. Manos. 2005. Molecular indicators of tree migration capacity under rapid climate change. Ecology 86: 2088–2098. McRae, B.H., Schumaker, N.H., McKane, R.B., Busing, R.T., Solomon, A.M and C.A. Burdick. 2008. A multi-model framework for simulating wildlife population response to land-use and climate change. Ecol. Mod. 219:77–91. Meffe, G.K. and C.R. Carroll. 1997. Principles of Conservation Biology, 2nd ed. Sinauer Associate, Inc. Sunderland, MA. Melles, S.J., Fortin, M.-J., Lindsay, K. and D. Badzinski. 2011. Expanding northward: Influence of climate change, forest connectivity, and population processes on a ’ range shift. Global Change Biol. 17:17–31. Menéndez, R., González-Megías, A.G., Collingham, Y., Fox, R., Roy, D.B., Ohlemüller, R. and C.D. Thomas. 2007. Direct and indirect effects of climate and habitat factors on butterfly diversity. Ecology 88: 605–611. Menéndez, R., Gonzalez-Megias, A., Lewis, O.T., Shaw, M.R. and C.D. Thomas. 2008. Escape from natural enemies during climate-driven range expansion: a case study. Ecol. Entomol. 33: 413–421. Menzel, A. and P. Fabrian. 1999. Growing season extended in Europe. Nature 397: 659. Millar, J.S. and E.J. Herdman. Climate change and the initiation of spring breeding by deer mice in the Kananaskis Valley, 1985-2003. Can. J. Zool. 82: 1444-1450. Miller, F.L. and A. Gunn. 2003. Catastrophic die-off of Peary caribou on the western Queen Elizabeth Islands, Canadian High Arctic. Arctic 56: 381-390. Miller, M.W., Swanson, H.M., Wolf, L.L., Quartarone, F.G., Huwer, S.L., Southwick, C.H. and P.M. Lukacs. 2008. Lions and prions and deer demise. PLOS One 3(2): e4019. Miller, W., et al. 2012. Polar and brown bear genomes reveal ancient admixture and demographic footprints of past climate change. Proc. Nat. Acad. Sci. USA 109:e2382–e2390. MNR Risk Management. 2013. Using adaptive management to deal with uncertainty in risk management. Integrating Risk in MNR, Tipsheet # 8. Møller, A.P., Fiedler, W. and P. Berthold. (eds.). 2004. Birds and Climate Change. Advances in Ecological Research, vol. 35. Elsevier Academic Press, London, UK. Møller, A.P., Fiedler, W. and P. Berthold. (eds.). 2010. Effects of Climate Change on Birds. Oxford Univ. Press, Oxford, UK. Mora, C., Metzker, R., Rollo, A. and R.A. Myers. 2007. Experimental simulations about the effects of habitat fragmentation and on populations facing environmental warming. Proc. R. Soc. Lond. B 274, 1023–1028. Morgan, E.R., Milner-Gulland, E.J., Torgerson, P.R. and G. F. Medley. 2004. Ruminating on complexity: macroparasites of wildlife and livestock. Trends Ecol. Evol. 19: 181–188. Moritz, C., Patton, J.L., Conroy, C.J., Parra, J.L., White, G.C. and S.R. Beissinger. 2008. Impact of a century of climate change on small-mammal communities in Yosemite National Park, USA. Science 322: 261–264. Morrison, L.W., Porter, S.D., Daniels, E. and M.D. Korzukhin. 2004. Potential global range expansion of the invasive fire ant, Solenopsis invicta. Biol. Invasions 6183–191. Muhlfeld, C.C., McMahon, T.E., Boyer, M.C. and R.E. Gresswell. 2009. Local-habitat, watershed, and biotic factors in the spread of hybridization between native westslope trout and introduced rainbow trout. Trans. Am. Fish, Soc. 138: 1036–1051. Murdoch, W.W., Briggs, C.J. and R.M. Nisbet. 2003. Consumer Resource Dynamics. Princeton University Press, Princeton, NJ. Murphy-Klassen, H.M., Underwood, T.J., Sealey, S.G. and A.A. Czyrnyj. 2005. Long-term trends in spring arrival dates of migrant birds at Delta marsh, Manitoba, in relation to climate change. Auk 122: 1130–1148. Murray, D.L., Boutin, S. and M. O’Donoghue. 1994. Winter habitat selection by lynx and coyotes in relation to snowshoe hare abundance. Can. J. Zool. 72: 1444–1451. Murray, D.L., Cox, E.W., Ballard, W.B., Whitlaw, H.A., Lenarz, M.E., Custer, T.W., Barnett, T. and T.K. Fuller. 2006. Pathogens, nutritional deficiency, and climate influences on a declining moose population. Wildl. Monogr. 166: 1–30. Myers, P., Lundrigan, B.L., Hoffman, S.M., Poor-Haraminac, A. and S.H. Seto. 2009. Climate induced changes in the small mammal communities of the northern Great Lakes region. Glob. Chang. Biol. 15: 1434–1454. Nicotra, A.B., Atkin, O.K., Bonser, S.P., Davidson, A.M., Finnegan, E.J., Mathesius, U., Poot, P., Purugganan, M.D., Richards, C.L., Valladares, F. and M. van Kleunen. 2010. Plant phenotypic plasticity in a changing climate. Trends Plant Sci. 15: 684–92. Nussey, D.H., Postma, E., Gienapp, P. and M.E. Visser. 2005. Selection on heritable phenotypic plasticity in a wild bird population. Science 310: 304–306. Obbard, M.E., Cattet, M.R.L., Moody, T., Walton, L.R., Potter, D., Inglis, J. and C. Chenier. 2006. Temporal trends in the body condition of Southern Hudson Bay polar bears. Ont. Min. Nat. Resour., Appl. Res. Devel. Br., Sault Ste. Marie, ON. Clim. Change Res. Inf. Note 3. Ogden, N.H., Maarouf, A., Barker, I.K., Bigras-Poulin, M., Lindsay, L.R., Morshed, M.G., O’Callaghan, C.J., Ramsay, F., Waltner-Toews, D. and D.F. CLIMATE CHANGE RESEARCH REPORT CCRR-36 31

Charron. 2006. Climate change and the potential for range expansion of the Lyme disease vector Ixodes scapularis in Canada. Int. J. Parasitol. 36: 63–70. Oosterkamp, T.E., Viereck, L., Shur, Y., Jorgenson, M.T., Racine, C., Doyle, A. and Boone, R.D. 2000. Observations of thermokarst and its impact in boreal forests in Alaska, U.S.A. Arctic Antarctic Alpine Res. 32: 303–315. Opdam, P. 1991. Metapopulation theory and habitat fragmentation: a review of holarctic breeding bird studies. Landscape Ecology 5: 93–106. Opdam, P. and D. Wascher. 2004. Climate change meets habitat fragmentation: linking landscape and biogeographical scale levels in research and conservation. Biol. Conserv. 117: 285–297. Opdam, P. and J.A. Wiens. 2002. Fragmentation, habitat loss and landscape management. Pp. 202–223 in Norris, K., Pain, D. (eds.). Conserving Bird Biodiversity. Cambridge University Press, Cambridge, UK. Paine, R.T. 1969. A note on trophic complexity and community stability. Am. Nat. 103: 91–93. Palmer, T.M., Stanton, M.L. and T.P. Young. 2003. Competition andcoexistence: exploring mechanisms that restrict and maintain diversity within mutualist guilds. Am. Nat.162:S63–S79 Parmesan, C. 1996. Climate and species’ range. Nature 382: 765–766. Parmesan, C. 2006. Ecological and evolutionary responses to recent climate change. Ann. Rev. Ecol. Evol. Syst. 37: 637–669. Parmesan, C. 2007. Influences of species, latitudes and methodologies on estimates of phenological response to global warming. Glob. Change Biol.13: 1860–1872. Parmesan, C. and G. Yohe. 2003. A globally coherent fingerprint of climate change impacts across natural systems. Nature 421: 37–42. Parmesan, C., et al.1999. Poleward shifts in geographical ranges of butterfly species associated with regional warming. Nature 399: 579–583. Patterson, B.R. and V.A. Power. 2002. Contributions of forage competition, harvest, and climate fluctuation to changes in population growth of northern white-tailed deer. Oecologia 130:62–71. Payette, S. 1987. Recent porcupine expansion at tree line: a dendroecological analysis. Can. J. Zool. 65: 551–557. Pearson, R.G. and T.P. Dawson. 2003. Predicting the impact of climate change on the distribution of species: are bioclimate envelope models useful? Global Ecol. Biogeogr. 12: 361–371. Peters, R.L. 1992. Conservation of biological diversity in the face of climate change. Pp. 59-71 in Peters, R.L. and T.E. Lovejoy (eds.). Global Warming and Biological Diversity. Yale University Press, New Haven, CT/London, UK. Peterson, G., Allen, C.R. and C.S. Holling. 1998. Ecological resilience, biodiversity and scale. Ecosystems 1: 6–18. Petit, J.R., et al. 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399: 429–436. Platt, T., Fuentes-Yaco, C. and K.T. Frank. 2003. Spring algal bloom and larval fish survival. Nature 423: 398–399. Pompe, S., Hanspach, J., Badeck, F., Klotz, S., Thuiller, W. and I. Kühn. 2008. Climate and land use change impacts on plant distributions in Germany. Biol. Lett. 4: 564–567. Post, E. 2003a. Large-scale climate synchronizes timing of flowering by multiple species. Ecology 84: 277–281. Post, E. 2003b. Timing of reproduction in large mammals. Pp. 437–450 in Schwartz M.D. (ed.). Phenology: An Integrative Environmental Science. Kluwer Academic Publ.,.New York, NY. Post, E. and M.C. Forchhammer. 2008. Climate change reduces reproductive success of an Arctic herbivore through trophic mismatch. Phil. Trans. R. Soc. Lond. B 363: 2367–2373. Post, E., C. Pedersen, C.C. Wilmers and M.C. Forchhammer. 2008. Warming, plant phenology, and the spatial dimension of trophic mismatch for large herbivores. Proc. R. Soc. Lond. B 275: 2005–2013. Post, E. and N.C. Stenseth. 1998. Large-scale climatic fluctuation and population dynamics of moose and white-tailed deer. J. Anim. Ecol. 67: 537–543. Post, E. and N.C. Stenseth. 1999. Climatic variability, plant phenology, and northern ungulates. Ecology 80: 1322–1339. Post, E., Peterson, R.O., Stenseth, N.C. and B.E. McLaren. 1999. Ecosystem consequences of wolf behavioural response to climate. Nature 401: 905–907. Pounds, J.A., et al. 2006. Widespread amphibian extinctions from epidemic disease driven by global warming. Nature 439: 161–167. Power, M.E., et al. 1996. Challenges in the quest for keystones. Bioscience 46: 609–620. Price, T.D., Qvarnström, A. and D.E. Irwin. 2003. The role of phenotypic plasticity in driving genetic evolution. Proc. R. Soc.Lond. B 270: 1433–1440. Pulido, F., Berthold, P. and A.J. van Noordwijk. 1996. Frequency of migrants and migratory activity are genetically correlated in a bird population: evolutionary implications. Proc. Nat. Acad. Sci. USA 93: 14642–14647. Qian, H. and R.E. Ricklefs. 2006. The role of exotic species in homogenizing the North American flora. Ecol. Lett. 9: 1293–1298. Quintero, I. and J. J. Wiens. 2013. Rates of projected climate change dramatically exceed past rates of climatic niche evolution among vertebrate species. Ecol. Lett.16: 1095–1103. Racey, G.D. 2004. Preparing for change: climate change and resource management in Northwest Region. Ont. Min. Nat. Resour., Northw. Sci. Info., Thunder Bay, ON. NWSI Tech. Workshop Rep. TWR-04. 32 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Réale, D., McAdam, A.G., Boutin, S. and D. Berteaux. 2003. Genetic and plastic responses of a northern mammal to climate change. Proc. R. Soc. Lond. B 270: 591–596. Régnière, J., Nealis, V. and K. Porter. 2009. Climate suitability and management of the gypsy moth invasion into Canada. Biol. Invasions 11: 135–148. Reh, W. and A. Seitz. 1990. The influence of land use on the genetic structure of populations of the common frog Rana temporaria. Biol. Conserv. 54: 239–249. Rejmánek, M. 1996. Species richness and resistance to invasions. Pp 153–172 in Orians, G., Dirzo, R., and J.H. Cushman (eds.). Diversity and Processes in Tropical Forest Ecosystems. Springer-Verlag, Berlin. Rejmánek, M. and D.M. Richardson. 1996. What attributes make some plant species more invasive? Ecology 77: 1655–1661. Rempel, R.S. 2012. Effects of climate change on moose populations : a vulnerability analysis for the Clay Belt ecodistrict (3E-1) in Northeastern Ontario. Ont. Min. Nat. Resour., Appl. Res. Devel. Br., Sault Ste. Marie, ON. Clim. Change Res. Rep. CCRR-26. 26 p. Rhymer, J.M. and D. Simberloff. 1996. Extinction by hybridization and introgression. Ann. Rev. Ecol. Syst. 27: 83–109. Richardson, D.M., et al. 2000. Invasive alien organisms and global change: a South African perspective. Pp. 303-349 in Mooney, H.A. and R.J.Hobbs (eds.). Invasive Speciesin a Changing World. Island Press, Washington, DC. Ricklefs, R.E. 1990. Ecology, 3rd ed. W.H. Freeman and Co., New York, NY. Ripple, W.J., Wirsing,A.J., Beschta, R.L. and S.W. Buskirk. 2011. Can restoring wolves aid in lynx recovery? Wildl. Soc. Bull. 35: 514–518. Ritchie, R.J. and R.E. Ambrose. 1996. Distribution and population status of bald eagles (Haliaeetus leucocephalus) in Interior Alaska. Arctic 49:120–128. Rizzo, B. and E. Wiken. 1992. Assessing sensitivity of Canada’s ecosystems to climatic change. Climatic Change 21:37–55. Root, T.L., Price, J.T., Hall, K.R., Schneider, S.H., Rosenzweig, C. and J.A. Pounds. 2003. Fingerprints of global warming on wild animals. Nature 421: 57–60. Rowley, I., Russell, E. and M. Brooker. 1993. In breeding in birds. Pp. 304-328 in Thornhill, N.W. (ed.).The Natural History of Inbreeding and Outbreeding. University of Chicago Press, Chicago. Roy, D.B. and T.H. Sparks. 2000. Phenology of British butterflies and climate change. Global Change Biol.6: 407–416. Rubidge, E.M., Monahan, W.B., Parra, J.L., Cameron, S.E. and J.S. Brashares. 2011. The role of climate, habitat, and species co-occurrence as drivers of change in small mammal distributions over the past century. Global Change Biol. 17: 696. Ruegg, K.C., Hijmans, R.J. and C. Moritz. 2006. Climate change and the origin of migratory pathways in the Swainson’s thrush, Catharus ustulatus. J. Biogeogr. 33: 1172–1182. Sala, O.E., et al. 2000. Global biodiversity scenarios for the year 2100. Science 287: 1770–1774. Sander, M., Balbao, T.C., Costa, E.S., dos Santos, C.R. and M.V. Petry. 2007. Decline in the breeding population of Pygoscelis antarctica and Pygoscelis adeliae on Penguin Island, South Shetland, Antarctica. Polar Biol. 30: 651−654 Saxon, E., Baker, B., Hargrove, W., Hoffman, F. and C. Zganjar. 2005. Mapping environments at risk under different global climate change scenarios. Ecol. Lett. 8: 53–60. Schiegg, K.,Pasinelli, G.,Walters, J.R and S.J. Daniels. 2002. Inbreeding and experience affect response to climate change in endangered woodpeckers. Proc. R. Soc. Lond. B. 269: 1153–1159. Schmitz O.J., Post, E., Burns, C.E. and K.M. Johnston. 2003. Ecosystem responses to global climate change: moving beyond color mapping. Bioscience 53: 1199–1205. Schweiger, O., et al. 2010. Multiple stressors on biotic interactions: how climate change and alien species interact to affect pollination. Biol. Rev. 85: 777–795. Schweiger, A.J., Lindsay, R.W., Vavrus, S. and J.A. Francis. 2008. Relationships between Arctic sea ice and clouds during autumn. J. Climate 21: 4799–4810. Scott, P.A. and I.T.M. Craine. 1993. The lynx cycle: a climatic perspective. Clim. Res. 2: 235–240. Selås, V. and J.O. Vik. 2007. The Arctic fox Alopex lagopus in Fennoscandia: a victim of human-induced changes in interspecific competition and predation? Biodivers. Conserv. 16: 3575–3583. Semlitsch, R.D. and J.R. Bodie. 1998. Are small, isolated wetlands expendable? Conserv. Biol. 12: 1129-1133. Settele, J., et al. 2008. Climatic Risk Atlas of European Butterflies. Pensoft, Sofia-Moscow. Sherry, R.A., Zhou, X., Gu, S., Arnone, J.A, III, Shimel, D.S, Verburg, P.S, Wallace, L.L. and Y. Luo. 2007. Divergence of reproductive phenology under climate warming. Proc. Nat. Acad. Sci. USA. 104: 198–202. Sillett, T.S., Holmes, R.T.and T.W. Sherry. 2000. Impacts of a global climate cycle on population dynamics of a migratory songbird. Science 288: 2040–2042. Sinclair, A.R.E., Fryxell, J.M. and G. Caughley. 2006. Wildlife Ecology, Conservation, and Management, 2nd ed. Blackwell Publishing, Oxford, UK. CLIMATE CHANGE RESEARCH REPORT CCRR-36 33

Smith, D. 1984. Ecology of the bobcat in a coniferous forest environment in western Montana 1984. Statewide Wildlife Research. Furbearing Mammal Studies. W-120-R-14 and 15 III FB-2.0 1. Sorvillo, F., Ash, L.R., Berlin, O.G.W., Yatabe, J., Degiorgio, C. and S.A. Morse. 2002. Baylisascaris procyonis: an emerging helminthic zoonosis. Emerging Infect. Dis. 8: 355–359. Spielman, D., Brook, B.W. and R. Frankham. 2004. Most species are not driven to extinction before genetic factors impact them. Proc. Nat. Acad. Sci. USA 101: 15261–15264. Stenseth,N.C., et al.1999. Common dynamic structure of Canada lynx populations within three climatic regions. Science 285: 1071–1073. Stenseth, N.C., et al. 2004. The effect of climatic forcing on population synchrony and genetic structuring of the Canadian lynx. Proc. Nat. Acad. Sci. USA 101: 6056–6061. Stenseth, N.C., Mysterud, A., Ottersen, G., Hurrell, J.W., Chan, K.-S. and M. Lima. 2002. Ecological effects of climate fluctuations. Science 297: 1292–1296. Stirling, I., Lunn, N.J. and Iacozza, J., Elliot, C. and M. Obbard. 2004. Polar bear distribution and abundance on the southwestern Hudson Bay coast during open water season, in relation to population trends and annual ice patterns. Arctic 57: 15–26. Stralberg, D., Jongsomjit, D., Howell, C.A., Snyder, M.A., Alexander, J.D., Wiens, J.A. and T.L. Root. 2009. Re-shuffling of species with climate disruption: A no-analog future for California birds? PLOS One 4(9):e6825. Strode, P.K. 2003. Implications of climate change for North American wood warblers (Parulidae). Global Change Biol. 9: 1137–1144. Swift, T.L. and S.J. Hannon. 2010. Critical thresholds associated with habitat loss: a review of the concepts, evidence, and applications. Biol.Rev. 85: 35–53. Thackeray, S.J., et al.. 2010. Trophic level asynchrony in rates of phenological change for marine, freshwater and terrestrial environments. Global Change Biol. 16: 3304–3313. Thomas, C.D., Bodsworth, E.J., Wilson, R.J., Simmons, A.D., Davies, Z.G., Musche, M. and L. Conradt. 2001. Ecological and evolutionary processes at expanding range margins. Nature 411: 577–581. Thomas, C.D., et al. 2004. Extinction risk from climate change. Nature 427: 145–148. Thomas, J.A., Rose, R.J., Clarke, R.T., Thomas, C.D. and N.R. Webb. 1999. Intra-specific variation in habitat availability among ectothermic animals near their climatic limits and their centers of range. Functional Ecol.13 (suppl. 1): 55–64. Thompson, I. and J. Baker. 2007. Climate change effects on caribou ecology. Pp. 8 in Rodgers, A.R., Allison, B.A., Wade, K.D. and E.P. Iwachewski (eds.). Forest-Dwelling Woodland Caribou in Ontario: Research Workshop Report. Ont. Min. Nat. Resour., Cent. North. For. Ecosys. Res., Thunder Bay, ON. Info. Pap. CNFER IP-001. 27pp. Thompson, I., Mackey, B. McNulty, S and A. Mosseler. 2009. Forest resilience, biodiversity, and climate change. A synthesis of the biodiversity/resilience/ stability relationship in forest ecosystems. Secretariat of the UN Convention on Biological Diversity, Montreal. Technical Series No. 43. Thuiller, W., et al. 2008. Predicting global change impacts on plant species’ distributions: future challenges. Perspect. Plant Ecol. Evol. Syst. 9: 137–52 Tilman, D. 1999. Ecological consequences of changes in biodiversity: a search for general principles. Ecology 80: 1455–1474. Tilman, D. and J.A. Downing. 1994. Biodiversity and stability in grasslands. Nature 367: 363–365. Torti, V.M. and P.O. Dunn. 2005. Variable effects of climate change on six species of North American birds. Oecologia 145: 486–495. Travis, J.M.J. 2003. Climate change and : a deadly anthropogenic cocktail. Proc. R. Soc. Lond. B 270: 467–473. Trivelpiece, W.Z., Trivelpiece, S.G. and N.J. Volkman. 1987. Ecological segregation of Adélie, gentoo, and chinstrap penguins at King George Island, Antarctica. Ecology 68: 351−361 Tylianakis J.M., Didham R.K., Bascompte J. and D.A. Wardle. 2008. Global change and species interactions in terrestrial ecosystems. Ecol. Lett. 11: 1351–1363. Tylianakis, J.M., Tscharntke, T. and O.T. Lewis. 2007. Habitat modification alters the structure of tropical host-parasitoid food webs. Nature 445: 202–205. Van der Putten, W.H., Macel, M. and M.E. Visser. 2010 Predicting species distribution and abundance responses to climate change: why it is essential to include biotic interactions across trophic levels. Phil. Trans. R. Soc. Lond. B 365: 2025–2034. van der Wal, R., Truscott, A.M., Pearce, I.S.K., Cole, L., Harris, M.P. and S. Wanless. 2008. Multiple anthropogenic changes cause biodiversity loss through plant invasion. Global Change Biol. 14: 1428–1436. Van Grunsven, R.H.A., Van der Putten, W.H., Bezemer T.M., Berendse, F. and E.M. Veenendaal. 2010. Plant–soil interactions in the expansion and native range of a poleward shifting plant species. Global Change Biol. 16: 380–385. van Noordwijk, A.J. and C.B. Müller. 1994. On adaptive plasticity in reproductive traits, illustrated with lay date in the great tit and colony inception in a bumble bee. Pp. 180–194 in: Jarman, P.J. and A. Rossiter (eds.). Animal societies; individuals, interactions and organization. Kyoto University Press, Kyoto. Varrin, R., Bowman, J. and P. A. Gray. 2007. The known and potential effects of climate change on biodiversity in Ontario’s terrestrial ecosystems: case studies and recommendations for adaptation. Ont. Min. Nat. Resour., Appl. Res. Devel. Br., Sault Ste. Marie, ON. Clim. Change Res. Rep. CCRR-09. 47 p. 34 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Vermeij, G.J. 1991. When biotas meet: understanding biotic interchange. Science 253: 1099–1104. Vila, M., et al. 2010. How well do we understand the impacts of alien species on ecosystem services? A pan-European cross-species assessment. Front. Ecol. Environ. 8: 135–144. Virkkala, R. and A. Rajasärkkä. 2011. Climate change affects populations of northern birds in boreal protected areas. Biol. Lett. 7: 395–398. Visser, M.E. 2008. Keeping up with a warming world; assessing the rate of adaptation to climate change. Proc. R. Soc. Lond. B 275: 649−659. Visser, M.E. and C. Both. 2005. Shifts in phenology due to global climate change: the need for a yardstick. Proc. R. Soc. Lond. B 272: 2561−2569. Visser, M. and J.M. Holleman. 2001. Warmer springs disrupt the synchrony of oak and winter moth phenology. Proc. R. Soc. Lond. B 268: 289−294. Visser, M.E, Both, C. and M.M. Lambrechts. 2004. Global climate change leads to mistimed avian reproduction. Adv. Ecol. Res. 35: 89–110. Visser, M.E., van Noordwijk, A.J., Tinbergen, J.M. and C.M. Lessells. 1998. Warmer springs lead to mistimed reproduction in great tits (Parus major). Proc. Roy. Soc. Lond. B. 265: 1867–1870. Voigt, D.R., Baker, J.A., Rempel, R.S. and I.D. Thompson. 2000. Forest vertebrate responses to landscape-level changes in Ontario. Pp. 198-233 in Perera, A.H., Euler, D.L. and Thompson, I.D. (eds.). Ecology of a Managed Terrestrial Landscape: Patterns and Processes of Forest Landscapes in Ontario. UBC Press, Vancouver, BC/Ont. Min. Nat. Resour., Toronto, ON. Waite, T.A. and D. Strickland. 2006. Climate change and the demographic demise of a hoarding bird living on the edge. Proc. R. Soc. Lond. B 273: 2809–2813. Walpole, A.A., Bowman, J., Tozer, D.C. and D.S. Badzinski. 2012. Community level response to climate change: shifts in anuran calling phenology. Herpetol. Conserv. Biol. 7: 249–257. Walther G.-R. 2010. Community and ecosystem responses to recent climate change. Phil. Trans. R. Soc. Lond. B 365: 2019–2024. Walther G.-R., et al. 2009. Alien species in a warmer world: risks and opportunities. Trends Ecol. Evol. 24: 686–693. Walther, G.-R., Post, E., Convey, P., Menzel, A., Parmesan, C., Beebee, T.J.C., Fromentin, J.-M.,Hoegh-Guldberg, O., and F. Bairllein. 2002. Ecological responses to recent climate change. Nature 416: 389–395. Warren, M.S., et al. 2001. Rapid responses of British butterflies to opposing forces of climate and habitat change. Nature 414: 65–69. Wasserman, T.N., Cushman, S.A., Shirk, A.S., Landguth, E.L., and J.S. Littell. 2012. Simulating the effects of climate change on population connectivity of American marten (Martes americana) in the northern Rocky Mountains, USA. Landscape Ecol. 27:211–225 Weigl, P.D. 1978. Resource overlap, interspecific interactions and the distribution of the flying squirrels, Glaucomys volans and G. sabrinus. Am. Midl. Nat. 100: 83–96. Weladji, R.B. and O. Holland. 2003. Global climate change and reindeer: effects of winter weather on the autumn weight and growth of calves. Oecologia 136: 317–323. Weller, W.F. 2009. Extension of the known range of the gray treefrog, Hyla versicolor, in northwestern Ontario. Can. Field-Nat. 123: 372. Wildlife Conservation Society. 2008. “Deadly Dozen” reports diseases worsened by climate change. Science Daily. http://www.sciencedaily.com/ releases/2008/10/081007073928.htm [accessed 2013 Mar 19] Williams, S.E., Shoo, L.P., Isaac, J.L., Hoffmann, A.A., and G. Langham. 2008. Towards an integrated framework for assessing the vulnerability of species to climate change. PLOS Biol 6: e325 Williams, B. K., Szaro, R. C. and C.D. Shapiro. 2009. Adaptive management. Adaptive Manage. Working Group, U.S. Dep. Interior., Washington, D.C.: U.S. Dep. Interior Tech. Guide. Wilmers, C.C., Sinha, S. and M. Brede. 2002. Examining the effects of species richness on community stability: an assembly model approach. Oikos 99: 363–367. Wilson, W.H., Kipervaser, D. and S.A. Lilley. 2000. Spring arrival dates of Maine migratory breeding birds: 1994-1997 vs. 1899-1911. Northeast. Nat. 7: 1–6. Winder, M. and D.E. Schindler. 2004 Climate change uncouples trophic interactions in an aquatic ecosystem. Ecology 85: 2100–2106. Wint, W. 1983. The role of alternative host plant species in the life of a polyphagous moth, Operophtera brumata (Geometridae). J. Anim. Ecol. 52: 439–450. Yang, L. H. and V.H.W. Rudolf. 2010. Phenology, ontogeny and the effects of climate changeon the timing of species interactions. Ecol. Lett. 13: 1–10. Yeh, P.J and T.D. Price. 2004. Adaptive phenotypic plasticity and the successful colonization of a novel environment. Am. Nat. 164: 531–542. Yom-Tov, Y. 2001. Global warming and body mass decline in Israeli passerine birds. Proc. R. Soc. Lond. B 268: 947–952. Yom-Tov, Y. and J. Yom-Tov. 2005. Global warming, Bergmann’s rule and body size in the masked shrew Sorex cinereus Kerr in Alaska. J. Anim. Ecol. 74: 803–808. Yurk B. P. and J.A. Powell. 2009. Modeling the evolution of insect phenology. Bull. Math. Biol. 71: 952–979. CLIMATE CHANGE RESEARCH REPORT CCRR-36 35

Appendix 1: Glossary

A glossary of technical terms used in this report. Adapted from Ayala (1982), IPCC (2007), Varrin et al. (2007), Ricklefs (1990), and Sinclair et al. (2006).

Asynchrony A discordance between or among processes.

Climate The average weather conditions of a defined area over a long period of time.

Climate envelope A description of the climate within which a species can persist; related to the fundamental niche.

Climate model A quantitative description of the interactions between the atmosphere, oceans, and land surface. Best guesses about these interactions are used to forecast how

changing CO2 levels will affect a future world. Community A group of interacting populations.

Demography The vital rates of a population; the study of the structure of a population.

Ecoregion A unique area nested within one of Ontario’s ecozones, defined by a characteristic range and pattern in climate, including temperature, precipitation, and humidity.

Hybridization Successful interbreeding between two different species, subspecies, or populations.

Keystone species Species that have a large environmental influence relative to their abundance.

North Atlantic Oscillation A north-south alternation in atmospheric mass that has large-scale effects on weather.

Phenology The study of the seasonality of animal and plant life.

Population A group of individuals of a single species in a particular area.

Refugia Locations of isolated or relict populations, where populations have persisted due to relatively benign conditions.

Regime shift Alteration to ecosystem services that have consequent impacts on human societies.

Species richness The number of different species in a defined area at a particular time.

Synergy The interaction of two processes such that the total effect is greater than each process acting independently.

Uncertainty An expression of the degree to which a value is unknown. It can result from lack of information or fromdisagreement about what is known or even knowable.

Weather The condition of the atmosphere over a short period of time, as described by various meteorological phenomena. 36 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Appendix 2: Summary of studies

A summary of studies of climate change effects on vertebrate species that occur in Ontario. Effects on the studied population(s) are noted as expansion, contraction, or equivocal. Studies include peer-reviewed journal articles or books in which long-term data (>5 years) were quantitively assessed for population responses to changing climate. Some additional studies were included if they were relevant to Ontario. Responses of vertebrate species that occur in Ontario but are not listed in the table were not found in the published literature. Of the 181 species listed in the table, reported effects are equivocal for 102, are consistent with expansion for 68, and are consistent with contraction for 11.

Documented effects of Common Class Scientific name climate Comment Sources name change on the population Spring call initiation Gibbs and Breisch Amphibia American Toad Bufo americanus EQUIVOCAL unchanged 1900-1912 to 2001 1990-1999 Spring call initiation earlier Gibbs and Breisch Amphibia Bullfrog Rana catesbeiana EXPANSION in 1990-1999 compared to 2001 1900-1912 Spring call initiation Amphibia Fowler’s Toad* Bufo fowleri EQUIVOCAL Blaustein et al. 2001 unchanged 1980 to 1998 Spring call initiation earlier in 1990-1999 compared to Gibbs and Breisch Amphibia Gray Treefrog Hyla versicolor EXPANSION 1900-1911, expansion into 2001, Weller 2009 Northern Ontario

Spring call initiation Gibbs and Breisch Amphibia Green Frog Rana clamitans EQUIVOCAL unchanged 1900-1912 to 2001 1990-1999

Northern Calling influenced by spring Amphibia Rana pipiens EXPANSION Walpole et al. 2012 Leopard Frog temperatures (1995-2008)

Leadback morph becoming Red-backed Gibbs and Karraker Amphibia Plethodon cinereus EQUIVOCAL more common, associated Salamander 2005 with warmer temperatures Blaustein et al. Breeds earlier in warmer Amphibia Spring Peeper Pseudacris crucifer EQUIVOCAL 2001, Gibbs and years Breisch 2001 Spring call initiation earlier in 1990-1999 compared to Gibbs and Breisch Amphibia Wood Frog Rana sylvatica EXPANSION 1900-1912, calling influenced 2001, Walpole 2012 by spring temperatures (1995-2008)

Temperature-dependent sex determination; grow larger Janzen 1994, Frazer Reptilia Painted Turtle Chrysemys picta CONTRACTION and reach maturity quicker et al. 1993 during warmer sets of years Spring arrival date became Alder Aves Empidonax alnorum CONTRACTION later 1899-1911 to 1994- Wilson et al. 2000 Flycatcher 1997 CLIMATE CHANGE RESEARCH REPORT CCRR-36 37

Spring arrival date in Maine unchanged 1899-1911 to Murphy-Klassen et American Aves Botaurus lentiginosus EQUIVOCAL 1994-1997; advancing with al. 2005, Wilson et Bittern warming temperatures in al. 2000 Manitoba Spring arrival is advancing Murphy-Klassen et Aves American Coot Fulica americana EXPANSION with warming temperatures al. 2005 in Manitoba

American Spring arrival is unrelated to Murphy-Klassen et Aves Falco sparverius EQUIVOCAL Kestrel temperature in Manitoba al. 2005

Spring arrival is later in Murphy-Klassen et American Maine; advancing with Aves Setophaga ruticilla EQUIVOCAL al. 2005, Wilson et Redstart warming temperatures in al. 2000 Manitoba Spring arrival is earlier in Murphy-Klassen et parts of its range, not in al. 2005, Torti and others; lays eggs earlier American Dunn 2005,Inouye Aves Turdus migratorius EQUIVOCAL in warmer springs; spring Robin et al. 2000, Wilson arrival is advancing with et al. 2000, Bradley warming temperatures in et al. 1999 Manitoba Spring arrival is earlier in Butler 2003, Wilson American Aves Scolopax minor EXPANSION parts of its range; calling et al. 2000, Bradley Woodcock earlier et al. 1999 Haliaeetus Documented population Ritchie and Ambrose Aves Bald Eagle* EXPANSION leucocephalus increase 1996 Spring arrival date unchanged 1899-1911 Murphy-Klassen et Baltimore Aves Icterus galbula EQUIVOCAL to 1994-1997 in Maine; al. 2005, Wilson et Oriole advancing with warming al. 2000 temperatures in Manitoba Murphy-Klassen Spring arrival is earlier in et al. 2005, Butler Aves Bank Swallow Riparia riparia EQUIVOCAL some parts of its range, 2003, Wilson et al. unchanged in others 2000 Spring arrival is earlier in some parts of its range, later Butler 2003, Wilson Aves Barn Swallow* Hirundo rustica EQUIVOCAL or unchanged in others; et al. 2000 clutch size increase Spring arrival is earlier in Bay-breasted Butler 2003, Wilson Aves Dendroica castanea EQUIVOCAL some parts of its range, Warbler et al. 2000 unchanged in others Spring arrival is earlier in Murphy-Klassen et some parts of its range, Belted al. 2005, Wilson et Aves Ceryle alcyon EQUIVOCAL unchanged in others; Kingfisher al. 2000, Bradley et advancing with warming al. 1999, temperatures in Manitoba Spring arrival date Black-and- Aves Mniotilta varia EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 white Warbler 1994-1997 Spring arrival became later 1899-1911 to 1994-1997,276 Wilson et al. 2000, Black-billed Coccyzus Aves EQUIVOCAL km northward range Hitch and Leberg Cuckoo erythropthalmus expansion 1967-1971 to 2007 1998-2002 38 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Spring arrival is earlier in Blackburnian Butler 2003, Wilson Aves Dendroica fusca EQUIVOCAL some parts of its range, later Warbler et al. 2000 in others Hybridization with Carolina Black-capped Aves Poecile atricapillus CONTRACTION chickadees whose range is Curry 2005 Chickadee expanding

Black-crowned Spring arrival in Manitoba is Murphy-Klassen et Aves Nycticorax nycticorax EQUIVOCAL Night-Heron unrelated to temperature al. 2005

Spring arrival date Blackpoll Aves Dendroica striata EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Warbler 1994-1997 Spring arrival date Black-throated Dendroica Aves EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Blue Warbler caerulescens 1994-1997

Spring arrival date Black-throated Aves Dendroica virens EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Green Warbler 1994-1997

314 kmnorthward range Blue-gray Hitch and Leberg Aves Polioptila caerulea EXPANSION expansion 1967-1971 to Gnatcatcher 2007 1998-2002

Spring arrival date Blue-headed Aves Vireo solitarius EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Vireo 1994-1997 Spring arrival is earlier in some parts of its range; Blue-winged Hitch and Leberg Aves Vermivora pinus EXPANSION 85 km northward range Warbler 2007, Butler 2003 expansion 1967-1971 to 1998-2002 Spring arrival is earlier in Murphy-Klassen some parts of its range, et al. 2005, Butler Aves Bobolink* Dolichonyx oryzivorus EQUIVOCAL later in others; unrelated to 2003, Wilson et al. warming temperatures in 2000 Manitoba Spring arrival date Broad-winged Aves Buteo platypterus EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Hawk 1994-1997 Spring arrival is unrelated Murphy-Klassen et Aves Brown Creeper Certhia familiaris EQUIVOCAL to warming temperatures in al. 2005 Manitoba Spring arrival is earlier in Murphy-Klassen parts of its range, not in et al. 2005, Butler Brown Aves Toxostomum rufum EQUIVOCAL others; advancing with 2003, Wilson et al. Thrasher warming temperatures in 2000, Bradley et al. Manitoba 1999 Spring arrival is unrelated Brown-headed Murphy-Klassen et Aves Molothrus ater EQUIVOCAL to warming temperatures in Cowbird al. 2005 Manitoba Onset of nesting earlier; Murphy-Klassen et spring arrival in Manitoba Aves Canada Goose Branta canadensis EXPANSION al. 2005, MacInnes is advancing with warming et al. 1990 temperatures Spring arrival date Canada Aves Wilsonia canadensis EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Warbler* 1994-1997 CLIMATE CHANGE RESEARCH REPORT CCRR-36 39

Cape May Spring arrival is earlier in Aves Dendroica tigrina EXPANSION Butler 2003 Warbler some parts of its range Aves Cardinal Cardinaalis cardinalis EXPANSION Calling earlier Bradley et al. 1999 Spring arrival date Chestnut-sided Dendroica Aves EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Warbler pensylvanica 1994-1997 Spring arrival is earlier in Butler 2003, Wilson Aves Chimney Swift* Chaetura pelagica EQUIVOCAL some parts of its range, later et al. 2000 in others Spring arrival date Chipping Aves Spizella passerina EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Sparrow 1994-1997 Chuck-will’s- Caprimulgus No significant range shift Hitch and Leberg Aves EQUIVOCAL widow carolinensis 1967-1971 to 1998-2002 2007 Spring arrival is unrelated Clay-coloured Murphy-Klassen et Aves Spizella pallida EQUIVOCAL to warming temperatures in Sparrow al. 2005 Manitoba Spring arrival date Petrochelidon Aves Cliff Swallow EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 pyrrhonota 1994-1997 Spring arrival advancing with Common Murphy-Klassen et Aves Quiscalus quiscula EXPANSION warming temperatures in Grackle al. 2005 Manitoba

Spring arrival became earlier Aves Common Loon Gavia immer EXPANSION Wilson et al. 2000 1899-1911 to 1994-1997

Spring arrival became later Murphy-Klassen et Common 1899-1911 to 1994-1997 in Aves Chordeiles minor EQUIVOCAL al. 2005, Wilson et Nighthawk* Maine; unrelated to warming al. 2000 temperatures in Manitoba

Spring arrival is earlier in Murphy-Klassen some parts of its range, not Common et al. 2005, Butler Aves Gallinago gallinago EQUIVOCAL in others; advancing with Snipe 2003, Wilson et al. warming temperatures in 2000 Manitoba Spring arrival date unchanged 1899-1911 Murphy-Klassen et Common Aves Geothlypis trichas EQUIVOCAL to 1994-1997 in Maine; al. 2005, Wilson et Yellow-throat unrelated to warming al. 2000 temperatures in Manitoba

Spring arrival in Manitoba is Murphy-Klassen et Aves Cooper’s Hawk Accipiter cooperii EXPANSION unrelated to temperature al. 2005

Spring arrival is advancing Dark-eyed Murphy-Klassen et Aves Junco hyemalis EXPANSION with warming temperatures Junco al. 2005 in Manitoba Spring arrival in Manitoba Double-crested Murphy-Klassen et Aves Phalacrocorax auritus EXPANSION is advancing with warming Cormorant al. 2005 temperatures Torti and Dunn Spring arrival is earlier in 2005, Butler Eastern some parts of its range, later Aves Sialia sialis EQUIVOCAL 2003,Wilson et al. Bluebird in Maine; lays eggs 4 days 2000, Bradley et al. earlier than in the 1970s 1999 40 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Spring arrival date unchanged 1899-1911 Murphy-Klassen et Eastern Aves Tyrannus tyrannus EQUIVOCAL to 1994-1997 in Maine; al. 2005, Wilson et Kingbird unrelated to warming al. 2000 temperatures in Manitoba Eastern Spring arrival is earlier in Aves Sturnella magna EXPANSION Bradley et al. 1999 Meadowlark* some parts of its range Spring arrival is earlier in Murphy-Klassen some parts of its range, Eastern et al. 2005, Butler Aves Sayornis phoebe EQUIVOCAL later in others; unrelated to Phoebe 2003, Bradley et al. warming temperatures in 1999 Manitoba Spring arrival is earlier in Eastern Wood- Butler 2003, Wilson Aves Contopus virens EQUIVOCAL some parts of its range, pewee et al. 2000 unchanged in others Spring arrival earlier in some Butler 2003, Wilson Aves Field Sparrow Spizella pusilla EQUIVOCAL parts of its range, later in et al. 2000 others Spring arrival earlier in parts Murphy-Klassen of its range, unchanged et al. 2005, Butler Aves Fox Sparrow Passerella iliaca EQUIVOCAL in others; advancing with 2003, Wilson et al. warming temperatures in 2000 Manitoba

Spring arrival earlier in some Golden-winged parts of its range, 148 km Butler 2003, Hitch Aves Vermivora chrysoptera EXPANSION Warbler* northward range expansion and Leberg 2007 1967-1971 to 1998-2002

Murphy-Klassen Spring arrival is earlier in et al. 2005, Butler Aves Gray Catbird Dumetella carolinensis EQUIVOCAL some parts of its range, 2003, Wilson et al. unchanged in others 2000 Populations decline following Waite and Strickland Aves Gray Jay Perisoreus canadensis CONTRACTION warmer autumns possibly 2006 due to hoard rot Gray-cheeked Spring arrival is earlier in Aves Catharus minimus EXPANSION Butler 2003 Thrush some parts of its range Spring arrival is earlier in Murphy-Klassen et Great Blue some parts of its range; al. 2005, Wilson et Aves Ardea herodias EXPANSION Heron advancing with warming al. 2000, Bradley et temperatures in Manitoba al. 1999 Spring arrival is earlier in Great Crested Butler 2003, Wilson Aves Myiarchus crinitus EQUIVOCAL some parts of its range, Flycatcher et al. 2000 unchanged in others Spring arrival is unrelated Greater Murphy-Klassen et Aves Tringa melanoleuca EQUIVOCAL to warming temperatures in Yellowlegs al. 2005 Manitoba Spring arrival is earlier in Aves Green Heron Butorides virescens EXPANSION Butler 2003 some parts of its range Henslow’s Ammodramus Spring arrival is earlier in Aves EXPANSION Butler 2003 Sparrow* henslowii some parts of its range Spring arrival is earlier in Murphy-Klassen some parts of its range, later et al. 2005, Butler Aves Hermit Thrush Catharus guttatus EQUIVOCAL in others; advancing with 2003, Wilson et al. warming temperatures in 2000 Manitoba CLIMATE CHANGE RESEARCH REPORT CCRR-36 41

Hooded 115 km range expansion Hitch and Leberg Aves Wilsonia citrina EXPANSION Warbler* 1967-1971 to 1998-2002 2007 Spring arrival is advancing Murphy-Klassen et Aves Horned Grebe* Podiceps auritus EXPANSION with warming temperatures al. 2005 in Manitoba Spring arrival is unrelated to Murphy-Klassen et Aves Horned Lark Eremophila alpestris EQUIVOCAL temperature in Manitoba al. 2005 Spring arrival is earlier in Murphy-Klassen some parts of its range; et al. 2005, Butler Aves House Wren Troglodytes aedon EQUIVOCAL unrelated to warming 2003, Bradley et al. temperatures in Manitoba 1999

Observed population Aves Ivory Gull Pagophila eburnea CONTRACTION Mallory et al. 2003 declines, Reduced sea ice

Spring arrival earlier in some Butler 2003, Wilson Aves Indigo Bunting Passerina cyanea EXPANSION parts of its range et al. 2000

148 km northward range Kentucky Hitch and Leberg Aves Oporornis formosus EXPANSION expansion 1967-1971 to Warbler 2007 1998-2002

Murphy-Klassen et Spring arrival is earlier in al. 2005, Torti and Aves Killdeer Charadrius vociferous EXPANSION some parts of its range; lays Dunn 2005, Butler earlier in warmer springs 2003

Aves Kirtland’s Setophaga kirtlandii CONTRACTION Habitat loss Botkin et al. 1991 Warbler*

Spring arrival date became later 1899-1911 to 1994- Murphy-Klassen et Least Aves Empidonax minimus EQUIVOCAL 1997 in Maine; unrelated to al. 2005, Wilson et Flycatcher advancing temperature in al. 2000 Manitoba Least Spring arrival is earlier in Aves Calidris minutilla EXPANSION Butler 2003 Sandpiper some parts of its range Spring arrival is unrelated Lesser Murphy-Klassen et Aves Tringa flavipes EQUIVOCAL to warming temperatures in Yellowlegs al. 2005 Manitoba Lincoln’s Spring arrival is earlier in Aves Melospiza lincolnii EXPANSION Butler 2003 Sparrow some parts of its range

Spring arrival is earlier in Louisiana parts of its range and later in Hitch and Leberg Aves Seiurus motacilla EQUIVOCAL Water-thrush* others; No significant range 2007, Butler 2003 shift 1967-1971 to 1998-2002

Spring arrival is earlier in Magnolia Butler 2003, Wilson Aves Dendroica magnolia EQUIVOCAL some parts of its range, later Warbler et al. 2000 in others Spring arrival is earlier in some parts of its range; Murphy-Klassen et Aves Marsh Wren Cistothorus palustris EQUIVOCAL unrelated to warming al. 2005, Butler 2003 temperatures in Manitoba 42 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Spring arrival is unrelated Murphy-Klassen et Aves Mourning Dove Zenaida macroura EQUIVOCAL to warming temperatures in al. 2005 Manitoba Spring arrival is earlier in Mourning Butler 2003, Wilson Aves Oporornis philadelphia EQUIVOCAL parts of its range and later or Warbler et al. 2000 unchanged in others Spring arrival is earlier in Nashville Butler 2003, Wilson Aves Vermivora ruficapilla EQUIVOCAL some parts of its range, Warbler et al. 2000 unchanged in others Spring arrival became later 1899-1911 to 1994-1997 Murphy-Klassen et Northern Aves Colaptes auratus EQUIVOCAL in Maine; advancing with al. 2005, Wilson et Flicker warming temperatures in al. 2000 Manitoba Spring arrival in Manitoba Northern Murphy-Klassen et Aves Circus cyaneus EXPANSION is advancing with warming Harrier al. 2005 temperatures Northern No significant range shift Hitch and Leberg Aves Mimus polyglottus EQUIVOCAL Mocking-bird 1967-1971 to 1998-2002 2007 Spring arrival is later in some Northern Butler 2003, Wilson Aves Parula americana EQUIVOCAL parts of its range, not in Parula et al. 2000 others Northern Stelgidopteryx Spring arrival is earlier in Aves Rough-winged EXPANSION Butler 2003 serripennis some parts of its range Swallow

Spring arrival is earlier in Northern Seiurus Butler 2003, Wilson Aves EQUIVOCAL some parts of its range, Water-thrush noveboracensis et al. 2000 unchanged in others Spring arrival date Olive-sided Aves Contopus cooperi EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Flycatcher* 1994-1997 Spring arrival is earlier in Aves Osprey Pandion haliaetus EXPANSION Butler 2003 some parts of its range Spring arrival date Aves Ovenbird Seiurus aurocapillus EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 1994-1997 Spring arrival is later in some parts of its range; unrelated Murphy-Klassen et Aves Palm Warbler Dendroica palmarum EQUIVOCAL to warming temperatures in al. 2005, Butler 2003 Manitoba Pectoral Spring arrival is earlier in Aves Calidris melanotos EXPANSION Butler 2003 Sandpiper some parts of its range Spring arrival date Philadel-phia Aves Vireo philadelphicus EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Vireo 1994-1997 Spring arrival in Manitoba Pied-billed Murphy-Klassen et Aves Podilymbus podiceps EXPANSION is advancing with warming Grebe al. 2005 temperatures Spring arrival date Aves Pine Warbler Dendroica pinus EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 1994-1997 No significant range shift Hitch and Leberg Aves Prairie Warbler Dendroica discolor EQUIVOCAL 1967-1971 to 1998-2002 2007 Spring arrival advancing with Murphy-Klassen et Aves Purple Finch Carpodacus purpureus EXPANSION warming temperatures in al. 2005 Manitoba CLIMATE CHANGE RESEARCH REPORT CCRR-36 43

Spring arrival is earlier in Murphy-Klassen some parts of its range, et al. 2005, Butler Aves Purple Martin Progne subis EQUIVOCAL unchanged in others; 2003, Wilson et al. unrelated to warming 2000 temperatures in Manitoba Spring arrival earlier in parts Red-eyed Butler 2003, Wilson Aves Vireo olivaceus EQUIVOCAL of its range and later or Vireo et al. 2000 unchanged in others Red-tailed Spring arrival in Manitoba is Murphy-Klassen et Aves Buteo jamaicensis EXPANSION Hawk unrelated to temperature al. 2005 Spring arrival is earlier in Murphy-Klassen et parts of its range; lays eggs al. 2005, Torti and Red-winged Aves Agelaius phoeniceus EXPANSION 7.5 days earlier; advancing Dunn 2005, Wilson Blackbird with warming temperatures et al. 2000, Bradley in Manitoba et al. 1999 Butler 2003, Wilson Rose-breasted Spring arrival is earlier in Aves Pheuticus ludovicianus EXPANSION et al. 2000, Bradley Grosbeak some parts of its range et al. 1999 Spring arrival is earlier in Murphy-Klassen some parts of its range, Ruby-crowned et al. 2005, Butler Aves Regulus calendula EQUIVOCAL unchanged in others; Kinglet 2003, Wilson et al. advancing with warming 2000 temperatures in Manitoba

Ruby-throated Spring arrival is earlier in Butler 2003, Wilson Aves Archilochus colubris EXPANSION Humming-bird some parts of its range et al. 2000

Spring arrival date Rusty Aves Euphagus carolinus EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Blackbird* 1994-1997 Spring arrival is unrelated Murphy-Klassen et Aves Sandhill Crane Grus canadensis EQUIVOCAL to warming temperatures in al. 2005 Manitoba Spring arrival is earlier in Savannah Passerculus Butler 2003, Wilson Aves EQUIVOCAL some parts of its range, later Sparrow sandwichensis et al. 2000 in others Spring arrival is earlier in Scarlet Butler 2003, Wilson Aves Piranga olivacea EQUIVOCAL some parts of its range, Tanager et al. 2000 unchanged in others Semi-palmated Spring arrival is earlier in Aves Calidris pusilla EXPANSION Butler 2003 Sandpiper some parts of its range

Sharp-shinned Spring arrival in Manitoba is Murphy-Klassen et Aves Accipiter striatus EQUIVOCAL Hawk unrelated to temperature al. 2005

Spring arrival is unrelated Short-eared Murphy-Klassen et Aves Asio flammeus EQUIVOCAL to warming temperatures in Owl* al. 2005 Manitoba Onset of nesting earlier; spring arrival in Manitoba Murphy-Klassen et unrelated to temperature, al. 2005, MacInnes Aves Snow Goose Chen caerulescens EQUIVOCAL reduction of some et al. 1990, Dickey components of breeding et al. 2003 success Solitary Spring arrival is earlier in Aves Tringa solitaria EXPANSION Butler 2003 Sandpiper some parts of its range 44 CLIMATE CHANGE RESEARCH REPORT CCRR-36

Spring arrival became later Murphy-Klassen et 1899-1911 to 1994-1997 in Aves Song Sparrow Melospiza melodus EQUIVOCAL al. 2005, Wilson et Maine; unrelated to warming al. 2000 temperatures in Manitoba

Spring arrival is earlier in some parts of its range; Murphy-Klassen et Aves Sora Porzana carolina EQUIVOCAL unrelated to advancing al. 2005, Butler 2003 temperature in Manitoba Spring arrival is earlier in Spotted Aves Actitis macularia EXPANSION some parts of its range; Murphy-Klassen et Sandpiper unrelated to warming al. 2005, Butler 2003 temperatures in Manitoba

Summer 43 km northward range Aves Piranga rubra EXPANSION Hitch & Leberg 2007 Tanager expansion 1967-1971 to 1998-2002 Spring arrival date Swamp Aves Melospiza melodus EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Sparrow 1994-1997

141 km northward range Hitch & Leberg Swainson’s Aves Catharus ustulatus EXPANSION expansion 1967-1971 to 2007, Ruegg et al. Thrush 1998-2002 2006 Tennessee Spring arrival is earlier in Butler 2003, Wilson Aves Vermivora peregrina EXPANSION Warbler some parts of its range et al. 2000

Pipilio Wilson et al. 2000, Aves Towhee EQUIVOCAL Did not show earlier arrival erythrophthalamus Bradley et al. 1999

Spring arrival advancing with Murphy-Klassen et Aves Tree Sparrow Spizella arborea EXPANSION warming temperatures in al. 2005 Manitoba Spring arrival is earlier some parts of its range; Average Murphy-Klassen egg-laying date up to 9 days et al. 2005, Butler earlier across NA; not laying 2003, Hussell 2003, Aves Tree Swallow Tachycineta bicolor EQUIVOCAL earlier at Long Point, Ontario Wilson et al. 2000, where temperatures have Dunn and Winkler not increased; unrelated to 1999 warming temperatures in Manitoba Spring arrival is earlier in Aves Turkey Vulture Cathartes aura EXPANSION Butler 2003 some parts of its range Spring arrival date Aves Veery Catharus fuscescens EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 1994-1997 Spring arrival became later Murphy-Klassen et Vesper 1899-1911 to 1994-1997 in Aves Pooecetes gramineus EQUIVOCAL al. 2005, Wilson et Sparrow Maine; unrelated to warming al. 2000 temperatures in Manitoba Spring arrival is earlier in Aves Virginia Rail Rallus limicola EXPANSION Butler 2003 some parts of its range Spring arrival is earlier in Butler 2003, Wilson Aves Warbling Vireo Vireo gilvus EQUIVOCAL some parts of its range, et al. 2000 unchanged in others Spring arrival is earlier in Wilson et al. 2000, Aves Whip-poor-will* Caprimulgus vociferus EXPANSION some parts of its range, later Bradley et al. 1999 in others CLIMATE CHANGE RESEARCH REPORT CCRR-36 45

Spring arrival is earlier in White-crowned some parts of its range; Murphy-Klassen et Aves Zonotrichia leucophrys EQUIVOCAL Sparrow unrelated to warming al. 2005, Butler 2003 temperatures in Manitoba

Spring arrival is earlier in Murphy-Klassen White-throated some parts of its range; et al. 2005, Butler Aves Zonotrichia albicollis EQUIVOCAL Sparrow unrelated to warming 2003, Wilson et al. temperatures in Manitoba 2000

135 km northward range Willow Aves Empidonax trailii EXPANSION expansion 1967-1971 to Hitch &Leberg 2007 Flycatcher 1998-2002 Spring arrival is earlier in Murphy-Klassen some parts of its range, Wilson’s et al. 2005, Butler Aves Wilsonia pusilla EQUIVOCAL unchanged in others; Warbler 2003, Wilson et al. advancing with warming 2000 temperatures in Manitoba Spring arrival date Troglodytes Aves Winter Wren EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 troglodytes 1994-1997 Spring arrival date Aves Wood Duck Aix sponsa EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 1994-1997 Spring arrival is earlier in Bradley et al. 1999, Aves Wood Thrush Hylocichla mustelina EXPANSION some parts of its range Butler 2003 Yellow Palm Dendroica palmarum Spring arrival is earlier in Aves EXPANSION Butler 2003 Warbler hypochrysea some parts of its range Murphy-Klassen Spring arrival is earlier in et al. 2005, Butler Aves Yellow Warbler Dendroica petechia EQUIVOCAL some parts of its range, 2003, Wilson et al. unchanged in others 2000 Spring arrival date Yellow-bellied Aves Empidonax flaviventris EQUIVOCAL unchanged 1899-1911 to Wilson et al. 2000 Flycatcher 1994-1997 Spring arrival is earlier in Yellow-bellied Butler 2003, Wilson Aves Sphyrapicus varius EQUIVOCAL some parts of its range, Sapsucker et al. 2000 unchanged in others Yellow-billed Spring arrival is earlier in Aves Coccyzus americanus EXPANSION Butler 2003 Cuckoo some parts of its range

Yellow- No significant range shift Hitch and Leberg Aves Icteria virens EQUIVOCAL breasted Chat* 1967-1971 to 1998-2002 2007 Murphy-Klassen Spring arrival is earlier in Yellow-rumped et al. 2005, Butler Aves Dendroica coronata EQUIVOCAL some parts of its range, Warbler 2003, Wilson et al. unchanged in others 2000 Yellow-throated Spring arrival is earlier in Aves Vireo flavifrons EXPANSION Butler 2003 Vireo some parts of its range Competition and predation by Selås and Vik 2007, red foxes expanding north, Mammalia Arctic Fox Alopex lagopus CONTRACTION Hersteinsson and changes in prey abundance, Macdonald 1992 habitat loss Reduced body weight of Weladji and Holland Mammalia Caribou* Rangifer tarandus CONTRACTION calves 2003 46 CLIMATE CHANGE RESEARCH REPORT CCRR-36

No effect of climate on Millar and Herdman Peromyscus initiation of spring breeding Mammalia Deer Mouse EQUIVOCAL 2004, Myers et al. maniculatus 1985 to 2003, Range 2009 contraction Phenotypic plasticity in Mammalia Elk Cervus elaphus EQUIVOCAL calving date during a 30 year Nussey et al. 2005 study Documented range Carr et al. 2007a,b; Mammalia Fisher Martes pennanti EXPANSION expansion, related to snow Voigt et al. 2000 depth Increased pack size in years Mammalia Gray Wolf Canis lupus EQUIVOCAL Post et al.1999 with deeper snow

Documented range Mammalia Least Weasel Mustela nivalis EXPANSION Frey 1992 expansion into Great Plains

Little Brown Energetic limit for hibernation Humphries et al. Mammalia Myotis lucifugus EXPANSION Bat shifting north 2002 Lynx-hare cycle related to climate; 175 km contraction Stenseth et al. 1999, Mammalia Lynx Lynx canadensis CONTRACTION of southern range limit in Koen et al. 2014 Ontario Possible contraction in Krohn et al. 1995, Mammalia Marten Martes americana EQUIVOCAL response to expanding 1997 fishers

Increased body size since Yom-Tov and Yom- Mammalia Masked Shrew Sorex cinereus EXPANSION 1950; documented range Tov 2005, Frey 1992 expansion into Great Plains Meadow Documented range Mammalia Jumping Zapus hudsonius EXPANSION Frey 1992 expansion into Great Plains Mouse

Microtus Documented range Mammalia Meadow Vole EXPANSION Frey 1992 pennsylvanicus expansion into Great Plains

Mink-muskrat cycle related Mammalia Mink Neovison vison EQUIVOCAL Haydon et al. 2001 to climate Increased disease at Murray et al. 2006, southern range boundary; Mammalia Moose Alces alces CONTRACTION Post and Stenseth cumulative effects of weather 1998 on body condition Mink-muskrat cycle related Mammalia Muskrat Ondatra zibethicus EQUIVOCAL Haydon et al. 2001 to climate Range contracts in response Northern Flying to competition from Bowman et al. 2005, Mammalia Glaucomys sabrinus CONTRACTION Squirrel expanding southern flying Weigl 1978 squirrel populations Kanda 2005, Austad Documented range Mammalia Opossum Didelphis virginiana EXPANSION 1988, Myers et al. expansion 2009 Obbard et al. 2006, Decreasing body Derocher et al. condition and productivity, Mammalia Polar Bear* Ursus maritimus CONTRACTION 2004, Stirling et al. hybridization, population 2004, Hunter et al. declines 2010

Porcupines following warming associated Voigt et al. 2000, Mammalia Porcupine Erethizon dorsatum EXPANSION poleward shift in tree line; Payette 1987 expansion related to reduced winter severity CLIMATE CHANGE RESEARCH REPORT CCRR-36 47

Documented range Larivière 2004, Voigt Mammalia Raccoon Procyon lotor EXPANSION expansion; related to et al. 2000 reduced winter severity Selås and Vik 2007, Expanding north due to Mammalia Red Fox Vulpes vulpes EXPANSION Hersteinsson and temperatures Macdonald 1992 Onset of breeding advanced Tamiasciurus Mammalia Red Squirrel EXPANSION by 18 days over a 10-year Réale et al. 2003 hudsonicus study Snowshoe Lynx-hare cycle related to Mammalia Lepus americanus EQUIVOCAL Stenseth et al. 1999 Hare climate Southern Energetic bottleneck shifting Bowman et al. 2005, Mammalia Glaucomys volans EXPANSION Flying Squirrel north, but dynamic boundary Weigl 1978, Myers et al. 2009,

White-footed EXPANSION Mammalia Peromyscus leucopus Northward range expansion Myers et al. 2009 mouse Garroway Cumulative effects of snow and Broders White-tailed depth reduce body condition 2005,Patterson and Mammalia Odocoileus virginianus EXPANSION Deer and fecundity; winter severity Power 2002, Voigt causes range contraction et al. 2000, Post and Stenseth 1999 * Species-at-risk Climate Change Research Publication Series Reports CCRR-19 Eskelin, N., W. C. Parker, S.J. Colombo and P. Lu. 2011. Assessing Assisted Migration as a Climate Change Adaptation Strategy for CCRR-01 Wotton, M., K. Logan and R. McAlpine. 2005. Climate Change Ontario’s Forests: Project Overview and Bibliography. and the Future Fire Environment in Ontario: Fire Occurrence and Fire Management Impacts in Ontario Under a Changing Climate. CCRR-20 Stocks, B.J. and P.C. Ward. 2011. Climate Change, Carbon Sequestration, and Forest Fire Protection in the Canadian Boreal Zone. CCRR-02 Boivin, J., J.-N. Candau, J. Chen, S. Colombo and M. Ter- Mikaelian. 2005. The Ontario Ministry of Natural Resources Large-Scale CCRR-21 Chu, C. 2011. Potential Effects of Climate Change and Adaptive Forest Carbon Project: A Summary. Strategies for Lake Simcoe and the Wetlands and Streams within the Watershed. CCRR-03 Colombo, S.J., W.C. Parker, N. Luckai, Q. Dang and T. Cai. 2005. The Effects of Forest Management on Carbon Storage in Ontario’s Forests. CCRR-22 Walpole, A and J. Bowman. 2011. Wildlife Vulnerability to Climate Change: An Assessment for the Lake Simcoe Watershed. CCRR-04 Hunt, L.M. and J. Moore. 2006. The Potential Impacts of Climate Change on Recreational Fishing in Northern Ontario. CCRR-23 Evers, A.K., A.M. Gordon, P.A. Gray and W.I. Dunlop. 2012. Implications of a Potential Range Expansion of Invasive Earthworms in CCRR-05 Colombo, S.J., D.W. McKenney, K.M. Lawrence and P.A. Gray. Ontario’s Forested Ecosystems: A Preliminary Vulnerability Analysis. 2007. Climate Change Projections for Ontario: Practical Information for Policymakers and Planners. CCRR-24 Lalonde, R., J. Gleeson, P.A. Gray, A. Douglas, C. Blakemore and L. Ferguson. 2012. Climate Change Vulnerability Assessment and CCRR-06 Lemieux, C.J., D.J. Scott, P.A. Gray and R.G. Davis. 2007. Adaptation Options for Ontario’s Clay Belt – A Case Study. Climate Change and Ontario’s Provincial Parks: Towards an Adaptation Strategy. CCRR-25 Bowman, J. and C. Sadowski. 2012. Vulnerability of Furbearers in the Clay Belt to Climate Change. CCRR-07 Carter, T., W. Gunter, M. Lazorek and R. Craig. 2007. Geological Sequestration of Carbon Dioxide: A Technology Review and Analysis of CCRR-26 Rempel, R.S. 2012. Effects of Climate Change on Moose Opportunities in Ontario. Populations: A Vulnerability Analysis for the Clay Belt Ecodistrict (3E-1) in Northeastern Ontario. CCRR-08 Browne, S.A. and L.M Hunt. 2007. Climate Change and Nature- based Tourism, Outdoor Recreation, and Forestry in Ontario: Potential CCRR-27 Minns, C.K., B.J. Shuter and S. Fung. 2012. Regional Projections Effects and Adaptation Strategies. of Climate Change Effects on Ice Cover and Open-Water Duration for Ontario Lakes CCRR-09 Varrin, R. J. Bowman and P.A. Gray. 2007. The Known and Potential Effects of Climate Change on Biodiversity in Ontario’s Terrestrial CCRR-28 Lemieux, C.J., P. A. Gray, D.J. Scott, D.W. McKenney and S. Ecosystems: Case Studies and Recommendations for Adaptation. MacFarlane. 2012. Climate Change and the Lake Simcoe Watershed: A Vulnerability Assessment of Natural Heritage Areas and Nature-Based CCRR-11 Dove-Thompson, D. C. Lewis, P.A. Gray, C. Chu and W. Dunlop. Tourism. 2011. A Summary of the Effects of Climate Change on Ontario’s Aquatic Ecosystems. CCRR-29 Hunt, L.M. and B. Kolman. 2012. Selected Social Implications of Climate Change for Ontario’s Ecodistrict 3E-1 (The Clay Belt). CCRR-12 Colombo, S.J. 2008. Ontario’s Forests and Forestry in a Changing Climate. CCRR-30 Chu, C. and F. Fischer. 2012. Climate Change Vulnerability Assessment for Aquatic Ecosystems in the Clay Belt Ecodistrict (3E-1) of CCRR-13 Candau, J.-N. and R. Fleming. 2008. Forecasting the Response Northeastern Ontario. to Climate Change of the Major Natural Biotic Disturbance Regime in Ontario’s Forests: The Spruce Budworm. CCRR-31 Brinker, S. and C. Jones. 2012. The Vulnerability of Provincially (Species at Risk) to Climate Change in the Lake Simcoe CCRR-14 Minns, C.K., B.J. Shuter and J.L. McDermid. 2009. Regional Watershed, Ontario, Canada Projections of Climate Change Effects on Ontario Lake Trout (Salvelinus namaycush) Populations. CCRR-32 Parker, W.C., S. J. Colombo and M. Sharma. 2012. An Assessment of the Vulnerability of Forest Vegetation of Ontario’s Clay Belt CCRR-15 Subedi, N., M. Sharma, and J. Parton. 2009. An Evaluation of (Ecodistrict 3E-1) to Climate Change. Site Index Models for Young Black Spruce and Jack Pine Plantations in a Changing Climate. CCRR-33 Chen, J, S.J. Colombo, and M.T. Ter-Mikaelian. 2013. Carbon Stocks and Flows From Harvest to Disposal in Harvested Wood Products CCRR-16 McKenney, D.W., J.H. Pedlar, K. Lawrence, P.A. Gray, S.J. from Ontario and Canada. Colombo and W.J. Crins. 2010. Current and Projected Future Climatic Conditions for Ecoregions and Selected Natural Heritage Areas in Ontario. CCRR-34 J. McLaughlin, and K. Webster. 2013. Effects of a Changing Climate on Peatlands in Permafrost Zones: A Literature Review and CCRR-17 Hasnain, S.S., C.K. Minns and B.J. Shuter. 2010. Key Ecological Application to Ontario’s Far North. Temperature Metrics for Canadian Freshwater Fishes.

CCRR-18 Scoular, M., R. Suffling, D. Matthews, M. Gluck and P. Elkie. CCRR-35 Lafleur, B., N.J. Fenton and Y. Bergeron. 2013. The Potential 2010. Comparing Various Approaches for Estimating Fire Frequency: The Effects of Climate Change on the Growth and Development of Case of Quetico Provincial Park. Forested Peatlands in the Clay Belt (Ecodistrict 3E-1) of Northeastern Ontario. 52762 (0.2k P.R.,13 12 30) ISBN 978-1-4606-3216-1 (print) ISBN 978-1-4606-3217-8 (pdf)